Cross-coupling reactions catalyzed by transition metal complexes. Classic cross-match

Candidate of Chemical Sciences Olga Belokoneva.

Modern man needs more and more complex, sophisticated substances - new antibiotics, cancer drugs, antivirals, plant protection products, light-emitting molecules for microelectronics. The 2010 Nobel Prize recognized an achievement in organic chemistry that sparked a breakthrough in the chemical industry by providing a versatile tool for creating unique compounds with a given chemical structure.

Cross-coupling reaction on a palladium catalyst using the Negishi reaction as an example.

Richard F. Heck was born in Springfield (USA) in 1931 and received his degree from the University of California. Heck is currently an honorary professor at the University of Delaware (USA). US citizen.

Ei-ichi Negishi was born in 1935 in Changchun, China and received his degree from the University of Pennsylvania. Currently, he is an honorary professor at Purdue University (USA). Japanese citizen.

Akira Suzuki (Akira Suzuki) was born in 1930 in Mukawa (Japan), received a degree from the University of Hokkaido (Japan). Currently, he is an honorary professor at the same university. Japanese citizen.

Professor Negishi during a lecture at Purdue University after the announcement of his Nobel Prize.

Richard Heck lectures at the University of Delaware (late 1960s).

Akira Suzuki at the International Symposium at the Institute of Organic Chemistry RAS in Moscow, September 2010.

You have to love chemistry. This is a very beautiful science that describes the processes taking place in the world of atoms and molecules. Chemistry must be respected, because the chemical compounds created by scientists allowed man to create a civilization that is so unlike the world of wildlife. And to understand how the world around us works - clothes, building materials, roads, cars, computers - you need to know chemistry.

The more complex substances a person needed on the path of progress, the more complex the chemical reactions that led to their creation became. At first, chemists followed the path of trial and error, then they learned to predict the course of reactions and create optimal conditions for the synthesis of a particular product. That's when it became possible to synthesize complex substances with unusual and useful properties. Most of them are organic compounds.

All living organisms are made up of organic compounds. It is so arranged in nature that the “molecular skeleton” of absolutely all organic molecules is a more or less complex chain of carbon atoms interconnected. The carbon-carbon bond is perhaps the most important chemical bond for all life on earth.

The carbon atom, like all other atoms, is a positively charged nucleus surrounded by layers of electron clouds. But for chemists, only the outer layer is of interest, because it is with the outer clouds that transformations usually occur, which are called chemical reactions. In the process of a chemical reaction, an atom seeks to complete its outer electron layer so that eight electrons “spin” around the nucleus. By itself, the carbon atom has only four outer electrons, therefore, in chemical bonding with other atoms, it seeks to socialize four “foreign” clouds in order to achieve the coveted stable “eight”. So, in the simplest organic molecule - methane, the carbon atom jointly "owns" electrons with four hydrogen atoms.

Now imagine that we need to synthesize a very complex organic molecule, similar to the one found in nature. Natural substances often have useful properties - they emit light, have an antitumor, antibacterial, analgesic effect, and polymerize. And to establish their laboratory synthesis is a very tempting task. Protein molecules are synthesized by genetic engineering, but non-protein ones have to be “cooked” manually in a chemical laboratory, which is not so simple. Several small organic molecules serve as the building blocks of a future complex natural structure. How to make them interact with each other? After all, the carbon atom in an organic molecule is stable and does not intend to enter into any reactions with other atoms.

To “stir up” the carbon atom, to make it reactive, is a truly Nobel task. At the beginning of the century, Victor Grignard, the 1912 Nobel laureate, first found a way to make carbon more active - he bonded it to a magnesium atom, as a result of which carbon became unstable and "started searching" for another carbon atom to form a chemical bond with it. And in total, for the entire existence of the Nobel Prizes, five (!) Prizes in chemistry have been awarded for the development of synthesis methods that lead to the creation of a bond between two carbon atoms. In addition to Grignard, Otto Diels and Kurt Alder (1950), Herbert C. Brown and Georg Wittig (1979), Yves Chauvin ), Robert H. Grubbs and Richard R. Schrock (2005).

And finally, the 2010 Nobel Prize was also awarded for a new method for creating a carbon-carbon bond. The Nobel Committee awarded the prize to Richard F. Heck, Ei-ichi Negishi and Akira Suzuki "for their application of cross-coupling reactions using palladium catalysts in organic synthesis." Cross-coupling reactions are organic reactions in which a chemical bond is formed between two carbon atoms that are part of different molecules.

Prior to the "palladium era" ushered in by the work of the current laureates, organic chemists had to synthesize complex molecules from blocks in several steps. Due to the high activity of the reagents, such a number of side compounds were formed in the reactions that the yield of the final product turned out to be scanty. The use of palladium was a very successful way out. It turned out to be an ideal "meeting place" for carbon atoms. On a palladium atom, two carbon atoms are so close to each other that an interaction can begin between them. The reaction on palladium proceeds with a high yield of the desired product without undesirable side processes.

This year's Nobel laureates have developed techniques for two types of reactions involving palladium. In both reactions, two reagents interact - electrophilic (with a deficit of electron density) and nucleophilic (with an excess of electron density). A hydrocarbon molecule (R) always acts as an electrophilic agent, in which the terminal hydrogen atom is replaced by a halogen atom (X = chlorine, bromine, iodine). But nucleophilic agents differ - in one case (Scheme 1) an olefin molecule (a linear hydrocarbon with one double bond) is used, and in the other (Scheme 2) an organometallic compound (M = zinc, boron or tin) is used. First, a complex of the palladium atom with an electrophilic agent is formed, and then this complex interacts with a nucleophilic compound.

The very idea of ​​using transition metals, including palladium, in organic synthesis arose long before the work of the current Nobel laureates. In the 1950s in Germany, for the first time, a palladium catalyst began to be used for the industrial oxidation of ethylene to acetaldehyde (the Wacker process), an important raw material for the production of paints, plasticizers, and acetic acid.

At the time, Richard Heck was working for a chemical company in Delaware. He became interested in the Wacker process and started experimenting with palladium. In 1968, Heck published a series of scientific papers on organometallic synthesis using olefins. Among them is a new way of "crosslinking" a simple olefin molecule with a benzene ring. The product of this reaction is vinylbenzene, from which polystyrene plastic is obtained.

Four years later, he developed a new method using olefins, which today is called the Heck reaction. It was for this achievement that he was awarded the Nobel Prize. The innovation was not only in olefins, but also in the use of hydrocarbon compounds with halogens as electrophilic agents. With the help of the Heck reaction, today they receive: the anti-inflammatory drug naproxen (Naproxen), the asthma drug - Singulair (Singulair), light-emitting compounds for microelectronics, taxol (Taxol) - a common chemotherapy drug. In a not very trivial way - in several stages - this method makes it possible to obtain the natural drug morphine and its chemical modifications. The Heck reaction is also used for the synthesis of steroid hormones (sex hormones, hormones of the adrenal cortex) and strychnine.

In 1977, Eichi Negishi was the first to use a zinc compound as a nucleophilic agent instead of olefins. Such reagents do not give unnecessary by-products, the yield of the final product is very high. The Negishi reaction allowed chemists to "sew" together complex functional groups that were impossible to synthesize "according to Heck".

Two years later, Akira Suzuki first used a compound containing a boron atom as a nucleophile. The stability, high selectivity, and low reactivity of organic boron compounds have made the Suzuki reaction one of the most useful in terms of practical application in industrial production. Boron compounds have low toxicity, reactions with their participation proceed under mild conditions. All this is especially valuable when it comes to the production of tens of tons of a product, such as the fungicide Boscalid (Boscalid), a means of protecting crops from fungal diseases.

One of the impressive achievements of the Suzuki method was the 1994 synthesis of palatoxin, a natural poison found in Hawaiian corals. Palatoxin consists of 129 carbon atoms, 223 hydrogen atoms, three nitrogen atoms and 54 oxygen atoms. The synthesis of such a huge organic molecule has inspired other feats of chemists. The Suzuki reaction has become a powerful tool in the chemistry of natural compounds. Indeed, only by synthesizing an artificial analogue in a test tube and comparing its properties with a natural substance, one can reliably confirm the chemical structure of a particular natural compound.

Now the eyes of organic chemists are largely turned towards the oceans, which can be considered as a warehouse of pharmaceutical products. Marine life, or rather, the physiologically active substances that they secrete, today serve as the main source of progress in the creation of new drugs. And in this, the reactions of Negishi and Suzuki help scientists. So, chemists managed to synthesize dasonamide A from the Philippine ascidian, which showed itself well in the fight against bowel cancer. A synthetic analogue of dragmacidin F from a sea sponge from the Italian coast affects HIV and herpes. Discodermolide from the sea sponge of the Caribbean Sea, which is synthesized using the Negishi reaction, is very similar in functional activity to taxol.

Palladium catalysts help not only synthesize natural compounds in the laboratory, but also modify existing drugs. This happened with vancomycin, an antibiotic that has been used since the middle of the last century to treat Staphylococcus aureus. During the time that has passed since the beginning of the use of the drug, bacteria have acquired resistance to it. So now, with the help of palladium catalysis, more and more new chemical modifications of vancomycin have to be synthesized, which even resistant bacterial specimens can handle.

Organic molecules capable of emitting light are used in the production of LEDs. Such complex molecules are also synthesized using the Negishi and Suzuki reaction. Chemical modification of light-emitting molecules makes it possible to increase the intensity of the blue glow under the influence of an electric current. Organic Light Emitting Diodes (OLEDs) are used in the production of super-thin, only a few millimeters thick, displays. Such displays are already used in mobile phones, GPS-navigators, night vision devices.

Synthesis using a palladium catalyst is used in the pharmaceutical industry, the production of plant protection products, and high-tech materials. With the help of cross-coupling reactions, it is possible to create analogues of natural compounds of almost any molecular configuration, which is very important for understanding the relationship between the structure and properties of complex organic molecules.

The reactions of Heck, Suzuki and Negishi are constantly modified and supplemented by other chemists. One of these innovations is associated with this year's Nobel Prize in Physics. The scientists succeeded in attaching palladium atoms to the molecular lattice of graphene, and the resulting solid-supported catalyst was successfully used to carry out the Suzuki reaction in an aqueous medium. The practical use of graphene is a matter of the future, and cross-coupling reactions on a palladium catalyst have already done a great service to humanity, although in fact their triumphant march is just beginning.

First, let's look at the general patterns of cross-coupling reactions.

Base metals in cross-coupling chemistry

these are group 10 metals in oxidation state 0 (zerovalent metals). Coordination-unsaturated complexes enter into the reaction. Of the three metals, palladium is the most versatile, nickel has a much narrower use, and platinum has no use at all.

The most commonly used complexes are zerovalent metals

with some simple and readily available ligands: nickel bis-cyclooctadiene complex, palladium tetrakis(triphenylphosphine) complex, and palladium dibenzylideneacetone complex, which exists in several forms.

triflates

- a very important type of electrophiles, allowing the use of a huge amount of phenols and enolizable carbonyl compounds in cross-coupling. But triflates are limited to derivatives with sp2 carbon, while halogen derivatives can have any type of electrophilic carbon.

Oxidative addition of chlorine derivatives

requires special ligands, for example, trialkylphosphines with bulky substituents - tris(tert-butyl)phosphine, tricyclohexylphosphine. The effect of these ligands is associated not only with high donation, but also with steric volume, which promotes the formation of coordinatively unsaturated active complexes.

Remetalization

this is the main way to load a nucleophile into the metal coordination sphere in a classical cross-coupling. With derivatives of magnesium, zinc, tin and other electropositive metals, remetalization occurs easily and does not require additional activation.

Reductive elimination is accelerated by phosphine chelators

especially those in which the angle between the bonds of the phosphine centers with the metal (bite angle) is greater than the standard for square planar right angle complexes. One of the most popular ligands of this type is dppf .

Cross-coupling - catalytic process

The active complex of zerovalent mell spontaneously regenerates after reductive elimination and enters a new round of the catalytic cycle. In the diagrams, the stages of the catalytic cycle are arranged in a circle, placing the active metal complex at the beginning of the cycle, which should be considered the actual catalyst.

Classic cross-combination.

The Big Four Major Cross-Coupling Reactions: Suzuki-Miyaura reaction(cross-coupling with organoboron compounds), Stille or Kosugi-Migita-Stille reaction (cross-coupling with organotin compounds), Negishi reaction (cross-coupling with organozinc compounds), Kumada or Kumada-Tamao-Corrio-Murahashi reaction (cross- combination with organomagnesium compounds).

The catalytic cycle of the Suzuki-Miyaura reaction operates in two ways, depending on the remetalation step, which needs additional activation (facilitation) either through the formation of a four-coordinate boron anion (the more common route) or through additional exchange of the ligand for palladium. In both cases, the reaction requires a rigid base with a negative charge on oxygen. For the same reason, the reaction is very often carried out in the presence of water to provide the hydroxide ion.

The Big Four is joined by an extremely important method of cross-coupling with terminal acetylenes - the Sonogashira or Sonogashira-Hagihara reaction, in which, at least formally, not an organometallic compound is used as a nucleophile, but directly a nucleophile - an acetylenide ion obtained directly in the reaction mixture from terminal acetylene. In fact, this is not entirely true, and this method is also based on the remetalation reaction.

New cross-combination. 1995-…

All of these classic reactions were discovered as early as the 1960s and 70s, and until the early 1980s developed into powerful methods of organic synthesis, which made it possible to synthesize thousands of previously inaccessible organic compounds. But by the beginning of the 1980s, the development of this field practically stopped, since there was no serious understanding of how to control the reactivity of metal complexes and overcome various obstacles, for example, low reactivity in reductive elimination, which does not allow one to “get” the product from the coordination sphere of the metal , and so on. Only after a decade and a half of intensive work on the study of mechanisms, the creation of new ligands and complexes, it was possible to move things forward, and an incredibly rapid growth of this science began in the mid-1990s. Methods discovered and developed after this milestone can be called the New Cross-Coupling. A special place in this new chemistry is no longer occupied by C-C cross-couplings, but by methods of forming carbon-another atom bonds. First of all, C-N bonds, the reactions of formation of which are very often, but not quite successfully, called amination.

Possibility of forming a C-N bond

in the cross-coupling reaction has been known since the early 1980s, when, for example, the reaction of bromobenzenes with a tin derivative of amines (the Kosugi-Migita reaction) was discovered, which is completely analogous to the Stille reaction in C-C cross-coupling. But this reaction did not find any application at all, not only because of the meager possibilities, but also because of the unwillingness of synthetics to contact toxic tin compounds.

The main task is how to use the amines themselves in the reaction

that is, to switch from transmetalization to direct substitution of the ligand to load the nucleophile into the coordination sphere. This problem was solved, but the resulting complex turned out to be stable to reductive elimination. It was possible to launch the last stage only when a suitable base was nazden, deprotonating the coordinated amine. However, the first useful ligand used in this process, tris(o-tolyl)phosphine, did not provide an opportunity to expand the range of amines due to side reactions and low yields.

BINAP is the most effective ligand

for C-N cross-coupling of bromo derivatives and triflates with secondary and primary amines, it not only effectively protects against the most annoying side process - the reductive dehalogenation of the bromo derivative, but also helps to push the reaction product out of the coordination sphere due to a significant steric volume.

Basic C-N Cross-Coupling Technique

uses BINAP as ligand and sodium t-butoxide as base. This technique has made it possible to obtain thousands of previously hard-to-find dialkylaryl-, alkyldiaryl, diaryl, and triarylamines with an extraordinary variety of substituents. The discovery of this reaction - the Hartwig-Buchwald (Batchwold) reaction - became a real revolution in the synthesis of nitrogen-containing compounds.

Development of new ligands

for example, new high-donor phosphines, which effectively control the metal coordination sphere due to steric factors and secondary coordination centers, have solved most of these problems and developed new selective protocols using chlorine derivatives and tosylates in reactions with higher catalytic efficiency (more TON)? significantly expand the synthetic range of the method.

Use of amides in C-N cross-coupling

long considered impossible, not only because of the low nucleophilicity, but also because of the chelate binding to the metal, which suppresses reductive elimination. Only with the introduction of special ligands, primarily the trans-chelating XantPhos, was it possible to solve these problems and make primary amides a legitimate substrate for C-N cross-coupling.

In addition to C-N cross-coupling, which became the No. 1 tool for the synthesis of various nitrogen compounds - amines, amides, hydrazines, heterocyclic compounds, even nitro compounds, etc., cross-coupling reactions began to be used to create carbon bonds with almost all non-metals and metalloids, and even with some metals. Let us choose from this almost infinite variety (quite monotonous, however, since all cross-coupling reactions follow the same pattern, which I hope we have already learned to understand quite well) carbon-boron bonding reactions, primarily because with their help we radically expand the capabilities of the Suzuki-Miyaura reaction, the main method of C-C cross-coupling.

is a typical cross-coupling reaction using a standard catalytic cycle including oxidative addition, transmetalation and reductive elimination. As a nucleophile, a diborane derivative, bis(pinacolato)dibor, is usually used, from which only one half is used.

although indirect, but the direct combination of electrophile-electrophile requires the participation of an additional reagent, a reducing agent, that is, it cannot be a catalytic process, and besides, it suffers from a congenital defect - the formation of mixtures of cross- and homocoupling. If we first convert one of the electrophiles to a nucleophile using Miyaura borylation, then we can then use the elaborate Suzuki-Miyaura cross-coupling.

in combination with the Suzuki-Miyaura cross-coupling also achieves the same goal of coupling two aryl fragments from two halogen derivatives or triflates, but requires a sequence of separate reactions that cannot be combined in the “one flask” mode.

So far, we have not gone beyond group 10 when discussing cross-coupling reactions. This is consistent with the dominant role of palladium and the important but secondary role of nickel in carbon-carbon bonding reactions. Until now, no other element has been able to displace this pair from the C-C cross-combination. But as soon as we switch to the bonds of carbon with other elements, the hegemony of palladium and nickel ends. next to them, another giant of catalysis appears - copper, an element of group 11, the ground valence state of which Cu(1+) has the same d 10 configuration as Ni(0). It is not surprising that this element was able to participate in a very similar chemistry, albeit with its own, exceptionally peculiar specificity. Surprisingly, silver has not been seen in anything like this, and Ag(1+) is just a spitting image of Pd(0), if we take into account the electronic configuration.

Copper cross-combination - the oldest cross-combination

The ability of copper to induce reactions that we now call cross-coupling has been known for over a hundred years. The Ulman-Goldberg reaction (not Goldberg, as it is sometimes written, Fritz Ullmann is the husband of Irma Goldberg) was used throughout the 20th century for the synthesis of diaryl- and triarylamines, arylamides and other compounds. The reaction requires very harsh conditions and uses active finely divided copper as either a reactant or a catalyst.

Reactions of Gilman Cuprates with Halogen Derivatives

Also a typical cross-combination, only stoichiometric. This reaction has been known and widely used since the 1950s. The electrophilic reagent in this reaction enters the copper coordination sphere due to nucleophilic SN2 substitution. The hypothetical mechanism of this reaction thus includes a typical cross-coupling change in oxidation state by 2 with regeneration of the original valence state after reductive elimination.

In the two previous sections, using examples of hydrogenation and isomerization reactions, we considered the main features of the mechanism of reactions catalyzed by transition metal compounds. Homogeneous hydrogenation and isomerization are very important reactions (despite the fact that at present, for economic reasons, hydrogenation - with the exception of asymmetric - is always carried out under heterogeneous conditions on the metals themselves), however, the most important reactions in organic synthesis are those that lead to the formation of new carbon-carbon bonds. In this and the following sections, such reactions will be considered. Let's start with the cross-coupling reaction.

Cross-coupling in the general sense is called reactions

RX + R "Y à RR" + XY,

where R are the organic groups that pair as a result of the reaction. Especially often in the synthesis, the interaction of s-organometallic compounds RM with organic halogen derivatives RX, catalyzed by soluble compounds of transition metals, taken in a catalytic amount, is used.

The role of the transition metal is that it initially enters into an oxidative addition reaction with an organic halide, and the resulting product (an alkyl compound of the transition metal) then rapidly reacts with an s-organometallic reagent, forming the cross-coupling product RR'. The catalytic cycle in its simplest form is shown in Scheme 27.6.

Since the metal increases its positive valence by two units in the catalytic cycle, it can be assumed that complexes containing the metal in low oxidation states should act as cross-coupling catalysts. Indeed, such reactions are catalyzed by soluble complexes of zerovalent metals (Ni, Pd, etc.). If complexes of divalent metals are used as a catalyst, for example, (Et 3 P) 2 NiCl 2, then zero-valent metal compounds are still formed during the reaction, for example, by the remetalization reaction

L 2 M II X 2 + R-m à L 2 M II (R)X + mX

With subsequent reductive elimination:

L 2 M II (R)X à + RX

The reaction then proceeds according to the cycle depicted in Scheme 27.6 (n = 2), through the steps of oxidative addition to RX and reductive elimination of ML 2 from R'ml 2 r .

Compounds of lithium, magnesium, zinc, boron, tin, mercury and other non-transition metals, and such transition metal compounds containing metal-carbon s-bonds, can be introduced into the cross-coupling reaction.

The limitation of the reaction is manifested when it is used for the synthesis of dialkyls (when R and R' are alkyl groups), since the yield of the cross-coupling product is significantly reduced due to possible b-elimination reactions (see section 27.8.4.b), leading to the formation alkenes:

The role of b-elimination is more noticeable when an alkyl halide containing hydrogen atoms in the b-position is introduced into the reaction than when an alkyl metal R-m (R \u003d alkyl with a b-atom H) reacts, since in equation 27.7 the b-elimination step (reaction b) competes to form a cross-coupling product (reaction a), and in equation 27.6 b - elimination occurs before the formation of L n M (R) (R ') turning into a cross-coupling product. Because of this limitation, cross-coupling is commonly used to prepare aryl- and vinylalkyl compounds.

The following are some examples of the synthetic use of the cross-coupling reaction:

(E)-Alkenyl complexes of zirconium obtained by the reaction of alkynes with Cp 2 Zr(H)Cl react with alkyl halides in the presence of palladium catalysts to form isomerically pure (97%) dienes in good yields. Complex LXVIII is as good in terms of yield and stereoselectivity as alkenyl aluminum compounds (Chapter 19, Section 19.3) and has the advantage that oxygen functions, such as ether or ketone groups, are not affected during the reaction.

Another group of transition metal complexes used in the synthesis of alkenes includes p-allylic compounds of nickel and palladium halides. These reagents are good because they can be obtained by a number of methods and, in the absence of contact with atmospheric oxygen, can be stored for several weeks. For example, Ni(II) p-allyl complexes are readily prepared from nickel carbonyl by heating with substituted allyl halides in benzene, or from bis-(1,6-cyclooctadiene)nickel and allyl halides at -10°C. The complexes have a dimeric bridging structure.

In polar coordinating solvents, these complexes react with many organic halides to form substituted alkenes, for example:

The presence of such functional groups as OH, COOR, COR, etc. does not interfere with the reaction.

p-Allyl complexes readily react with external anionic nucleophiles to form allyl nucleophilic substitution products. Of particular importance is the reaction with carbanions, because. in this case, a new C-C bond is formed in the allyl position.

Application of chiral phosphine ligands. as in the case of hydrogenation (see Section 27.9.1.c), it allows the asymmetric synthesis of alkenes. For example, the cross-coupling of a-phenylethylmagnesium chloride with vinyl bromide, catalyzed by nickel complexes containing chiral ligands based on ferrocenylphosphines, gives 3-phenyl-butene-1 in an optically active form.

As in the case of hydrogenation, the enantiomeric excess depends on the structure of the chiral ligand, and in this case the optical yield is increased if the chiral ligand contains a -NMe 2 group, which is probably coordinated to magnesium. So, if in the ligand (LXIX) X = H, then the enantiomeric excess is only 4%, but if X = NMe 2, then the enantiomeric excess increases to 63%.

1. Introduction.

2. Literature review.

2.1. Cross-coupling mechanism catalyzed by palladium(O) complexes stabilized by monodentan phosphine ligands.

2.1.1. Pd°L4 as PdL2 precursor (L = PPh3).

2.1.2. Pd°(dba)2 + nL (n>2) as PdL2 precursor (L = monodentate phosphine ligand).

2.1.3. Pd°(OAc)2 + nL (n>3) (L - PPh3).

2.1.4. PdX2L2 (X = halide, L = PPh3).

2.2. Structure of arylpalladium(II) complexes obtained by oxidative addition to aryl halides/triflates.

2.2.1. TpaHC-Ar?dXL2 (X = halide, L = PPh3).

2.2.2. Dimer complexes? (X = halide,

2.2.3. Cationic complexes ////."az/e-ArPdl^S4^ (S = solvent,

2.2.4. Equilibrium between the neutral ArPdXL2 complex and the cationic ArPdL2S+ (X = halide, L = PPh3).

2.2.5. Five-coordinate anionic complexes: ArPdXXiL2"

X and Xi = halides, L = PPh3).

2.2.6. Neutral w/?aH6"-ArPd(OAc)L2 complexes (L = PPh3).

2.3. Reactions of nucleophiles with arylpalladium complexes (remetllation).

2.3.1. Cationic complexes ArPdL2S+ (L = PPh3).

2.3.2. Dimer complexes 2 (X = halide,

2.3.3. Complexes w^mc-ArPd(OAc)L2 (L = PPh.O-.

2.3.4. Trans-ArPhoXb2 complexes (X = halide, L = monophosphine).

2.3.5. Five-coordinate anionic complexes: ArPdXXiL^"

X and Xi = halides, L = PPb3).

2.4. Mechanism of the cross-coupling reaction catalyzed by palladium(O) complexes stabilized by bidentate phosphine ligands.

2.4.1. Pd^V-L-IOOL-L) - as a precursor for obtaining Pd°(L-L)

2.4.2. Pd°(dba)2 and L-L - as a precursor for obtaining Pd°(L-L)

L = diphosphine lignd).

2.4.3. Remetalization of z/Mc-ArPdX(L-L) complexes.

2.4.4. Reductive elimination from */MC-ArPdNu(L-L) complexes.

2.5. General ideas about the Begishi reaction.

2.5.1. Methods for polunation of organozinc compounds.

2.5.1.1 Remetalization.

2.5.1.2 Oxidative zinc coating.

2.5.1.3 Zn-halogen exchange.

2.5.1.4 Zn-hydrogen exchange.

2.5.1.5 Hydrozinconation.

2.5.2. Influence of the nature of the electrophile (RX).

2.5.3. Palladium or nickel catalysts and ligands.

2.6. Use of the Tsegishi reaction to obtain biaryls.

2.7. Recent advances in the field of obtaining biaryls by the reaction of cross-coupling.

3. Discussion of the results.

3.1. Synthesis of yans-zirconocenes involving preliminary catalytic arylation of halogen-substituted bridging ligands.

3.1.1. Synthesis of halogenated b?/c(indenyl)dimethylsilanes and similar compounds.

3.1.2. Palladium-catalyzed arylation of 4/7-halogen-substituted bms(indenyl)dimethylsilanes and similar compounds.

3.1.3. Synthesis of ansch-zirconocenes from ligands obtained by cross-coupling reaction involving halogen-substituted bridging ligands.

3.2. Study of palladium-catalyzed arylation of halogen-substituted zirconium and hafnium complexes.

3.2.1. Synthesis and study of the structure of halogen-substituted complexes of zirconium and hafnium.

3.2.2. Study of palladium-catalyzed Negishi arylation involving halogen-substituted zirconium and hafnium complexes.

3.2.3. Study of palladium-catalyzed Suzuki-Miyaura arylation involving bromo-substituted zirconium complexes and NaBPht.

4. Experimental part.

5. Conclusions.

6. Literature.

List of abbreviations

DME dimethoxyethane

THF, THF tetrahydrofuran

DMF dimethylformamide

NML N-methylpyrollidone

NMI N-methylimidazole

MTBE methyl tertiary butyl ether

S solvent, solvent

TMEDA М^К.М"-tetramethylethylenediamine

Hal halogen

Nu nucleophile dba dibenzylideneacetone

Wed cyclopentadiene

Wed* pentamethylcyclopentadiene

Tolil

Ac acetyl

RG propyl

Su cyclohexyl

Alk, Alkyl alkyl

OMOM MeOSNGO

Piv pivaloyl

COD 1,5-cyclo-octadiene n, p normal and iso t, tertiary c, sec secondary o ortho p para cyclo equivalent

TON turnover number is one of the definitions: the number of moles of a substrate that can be converted into a product by 1 mole of a catalyst before losing its activity.

TTP tri(o-tolyl)phosphine

TFP tri(2-furyl)phosphine

DPEphos bis(o,o"-diphenylphosphino)phenyl ether

Dppf 1, G-bis(diphenylphosphino)ferrocene

Dipp 1,3 -bis(isopropylphosphino)propane

Dppm 1.1 "-bis(diphenylphosphino)methane

Dppe 1,2-bis(diphenylphosphino)ethane

Dppp 1,3-bis(diphenylphosphino)propane

Dppb 1,4-bis(diphenylphosphino)butane

DIOP 2,3-O-isopropylidene-2,3-dihydroxy-1,4-bis(diphenylphosphino)butane

B1NAP 2,2"-bis(diphenylphosphino)-1, G-binaphthyl

S-PHOS 2-dicyclohexylphosphino-2",6"-dimethoxybiphenyl

DTBAH, DTBAL diisobutyl aluminum hydride

NMR nuclear magnetic resonance

J spin-spin coupling constant

Hz Hz br broadened s singlet d doublet dd doublet doublet dt doublet triplet dkv doublet quadruplet t triplet m multiplet

M molar, metal sq quadruplet y broadened ml milliliter μm, | jap micrometer g gram ml milliliter otteor. from the theory they say. mole mole mimole others other

Tbp. boiling point h h cat. number catalytic amount vol. volume

MAO methylallumoxane

HPLC high performance liquid chromatography

Recommended list of dissertations

  • Study of approaches to the synthesis and structure of new bis-indenyl ansa-zirconocenes 2007, candidate of chemical sciences Izmer, Vyacheslav Valerievich

  • Halogen-substituted cyclopentadienyl-amide complexes of titanium and zirconium with strained geometry and cross-coupling reactions with their participation 2011, Candidate of Chemical Sciences Uborsky, Dmitry Vadimovich

  • Synthesis and study of ANSA-zirconocenes containing 4-NR2-2-methylindenyl fragments 2008, Candidate of Chemical Sciences Nikulin, Mikhail Vladimirovich

  • Phosphonium salts based on sterically loaded phosphines: synthesis and application in the Suzuki and Sonogashira reactions 2010, candidate of chemical sciences Ermolaev, Vadim Vyacheslavovich

  • Palladium(II) complexes with 1,1`-bis(phosphino)ferrocenes. Effect of Substituents at Phosphorus Atoms on Spectral, Structural, and Catalytic Properties 2007, candidate of chemical sciences Vologdin, Nikolai Vladimirovich

Introduction to the thesis (part of the abstract) on the topic "The use of palladium-catalyzed cross-coupling reactions for the synthesis of substituted cyclopentadienyl and indenyl complexes of zirconium and hafnium"

The production of polyolefins is one of the fundamental processes of modern industry, and most of these polymers are obtained using traditional heterogeneous Ziegler-type catalysts. An alternative to these catalysts are homogeneous and heterogenized Ziegler-Natta systems based on cyclopentadienyl derivatives of titanium subgroup metals, which make it possible to obtain new grades of polymers with improved physicochemical, morphological, granulometric properties, and other important consumer characteristics. Obviously, the theoretical models for transition metal compounds are difficult enough to predict the exact properties of the corresponding catalytic systems using modern high-level theoretical calculations. Therefore, today and in the near future, apparently, there is no alternative to the experimental enumeration of the corresponding catalysts and the conditions under which they are tested. This fully applies to cyclopentadienyl complexes of metals of the titanium subgroup. Therefore, the creation of new effective methods of synthesis, and in particular high-performance synthesis, of these complexes is currently an important scientific and applied task.

It is known that catalysts based on racemic ansa-metallocenes containing dimethylsilyl-bms-indenyl ligands with methyl in position 2 and an aryl substituent in position 4 (complexes of type A), as well as analogous complexes of type B, have high activity and stereoselectivity in the polymerization of propylene. containing 2,5-dimethyl-3-arylcyclopenta[£]thienyl fragments.

The main method for the synthesis of type A ansa-zirconocenes is the reaction between the dilithium salt of the s/c-indenyl ligand with zirconium tetrachloride. In turn, b's (indenyl)dimethylsilanes are obtained by the reaction of 2 equivalents of the lithium salt of the corresponding indene with dimethyldichlorosilane. This synthetic approach is not without drawbacks. Since the proton in the indenyl fragment of the intermediate product of this reaction, i.e. indenyldimethylchlorosilane, which is more acidic than in the starting indene, then during the synthesis of the bridging ligand, a side reaction of metallation of the intermediate with the lithium salt of indene occurs. This leads to a decrease in the yield of the target product, as well as to the formation of a large amount of side polymer/oligomeric compounds.

Continuing the logic of retrosynthetic analysis, it should be noted that the synthesis of aryl-substituted indenes is required to obtain the corresponding bms(indel)dimethylslanes. Aryl-substituted indenes can be obtained by the multi-stage "malon" method from the corresponding benzyl halides containing a biphenyl fragment in their structure. According to this synthetic approach, the starting benzyl halide is first reacted with the sodium or potassium salt of diethylmethylmalopic ether. After saponification of the ester and subsequent decarboxylation of the resulting diacid, it is possible to obtain the corresponding substituted propionic acid. In the presence of AlCl, the acid chloride of this acid is cyclized to form the corresponding indanone-1. Further reduction of substituted indanones-1 with sodium borohydride in a tetrahydrofuran-methanol mixture, followed by acid-catalyzed dehydration of the reduction products, leads to the formation of the corresponding indenes. This method is of little use and is very labor-intensive in the synthesis of a large number of similar aryl-substituted indenes. This is due to the fact that, firstly, benzene halides, which are the initial substrates in this synthesis, are not readily available compounds, and most of them must first be obtained. Secondly, a single multi-stage "small-op" synthesis makes it possible to obtain only one required aryl-substituted indene, and therefore, to obtain a number of products of the same type, this multi-stage synthesis must be carried out several times.

An alternative approach involving palladium-catalyzed arylation of halogenated indenes and similar substrates is more promising. Having received the "parent" halogen-substituted indene once, we are able to synthesize various aryl-substituted indenes in one stage. Despite the undeniable advantages of this approach, it is necessary to note its certain disadvantages. For example, to obtain a number of aryl-substituted apsa complexes of type A (or B), it is necessary to obtain a number of corresponding bridging ligands, i.e. carry out the appropriate number of reactions between the salt of indene (or its cyclopeitathienyl analogue) and dimethylchlorosilane. Then, several reactions must be carried out to synthesize the metallocenes themselves. It is assumed that a more productive approach consists in the preliminary synthesis of one "parent" halogen-substituted b//c(indenyl)dimethylsilane, which can be further used as a substrate for catalytic cross-coupling involving various aryl organoelement derivatives. This would make it possible to obtain various bridging leagues in one stage, and then the corresponding Yansa-metallocenes. Therefore, one of the goals of this work is the synthesis of bromo-substituted bis(icdenyl)dimethylsilanes and similar compounds, and then the development of methods for the palladium-catalyzed arylation of such substrates to obtain various aryl-substituted bridging ligands.

It should be noted that the use of such substrates in the cross-coupling reaction may be associated with certain difficulties. This is due to two circumstances. First, silyl derivatives of indenes are not completely inert compounds in the presence of palladium catalysts. These compounds, which include olefin and allylsilyl fragments, are potential substrates for the Heck and Hiyama reactions, respectively. Second, the silicon-cyclopentadienyl bond in o'c(indenyl)dimethylsilanes is known to be very sensitive to alkalis and acids, especially in protic media. Therefore, rather strict restrictions were initially imposed on the conditions for the implementation of catalytic arylation. In particular, carrying out the reaction in the presence of bases in protic solvents, for example, water, was completely excluded. The use of strong bases, such as ArMgX, which are substrates in the Kumada reaction, was also unacceptable, since it could be accompanied by metalation of indenyl fragments and a decrease in the yield of target compounds.

Undoubtedly, a synthetic method involving a cross-matching reaction with the participation of halogen-containing bms(indenyl)dimethylsplanes will make it possible to significantly simplify the preparation of a number of the same type of aryl-substituted n-metallocenes based on them, since it allows the introduction of an aryl fragment at a relatively late stage of synthesis. Guided by the same considerations, it can be assumed that the successful use of the corresponding Apsa complex as a “mother” substrate would be the simplest and most convenient method for obtaining structures of this type. Here, it must be emphasized that the use of complexes as substrates for the cross-coupling reaction is even more problematic than the use of bis(indenpl)dimethylsilanes. First, zirconium complexes interact with organolithium and organomagnesium compounds to form compounds with Zt-C bonds. Secondly, zirconium complexes, by themselves, are compounds sensitive to traces of water and air, which significantly complicates the work from a methodological point of view. Nevertheless, another goal of this work was to develop methods for the synthesis of halogen-substituted /Dcyclopentadienyl complexes of zirconium (and hafnium) of various types, as well as the subsequent study of the possibility of using these compounds as substrates in palladium-catalyzed Negishi and Suzuki-Miyaura cross-coupling reactions. .

Due to the fact that the Negishi reaction with the participation of organozinc compounds was used as the main method of cross-coupling of halogen-substituted substrates, the literature review of the dissertation is mainly devoted to the description of this particular method.

2. Literature review

The following literature review consists of three main parts. The first part describes the results of studies on the mechanisms of palladium-catalyzed cross-coupling reactions (Scheme 1). The possibility of effective implementation of the cross-coupling reaction depends on various factors, such as the nature of the precatalyst, the nature of the substrates, the solvent, and various additives. Thus, the purpose of the first part of the literature review, in addition to describing the reaction mechanisms, was to consider these dependencies. The second part of the literature review is devoted to the Negishi reaction, which is a cross-coupling catalyzed by palladium or nickel complexes involving various organic electrophiles and organozinc compounds. The history of the discovery of this method is briefly described, as well as the main factors that can affect the yield of the product in the Negishi reaction, i.e., the nature of the precatalyst, the nature of the substrates and solvent used. Cross-coupling with the participation of organozinc compounds catalyzed by palladium or nickel complexes has wide synthetic possibilities, making it possible to obtain a large number of valuable organic products. Cross-coupling reactions in general, and the Negishi method in particular, are often used to form the C(sp2)-C(sp2) bond. Thus, the development of conditions for carrying out cross-coupling reactions made it possible to efficiently synthesize various biaryls, the preparation of which by alternative methods seemed to be a very difficult task. The Negishi reaction makes it possible to obtain biaryls of various nature under fairly mild conditions and in good yields. The third part of the literature review is devoted to describing the possibilities of the Negishi reaction for the synthesis of various compounds containing a biaryl moiety. Moreover, the structure of the presentation is such that the synthetic possibilities of this method are considered in comparison with other main protocols for cross-coupling reactions. This type of presentation was chosen due to the importance of choosing the conditions for carrying out the cross-coupling reaction in the synthesis of specific compounds. It should be noted that due to the huge amount of information on this topic and the limitations imposed on the volume of the dissertation, the third part of the literature review outlines only the main, most characteristic features of the Negishi method. Thus, the topic of obtaining biaryls, in which one or both aryl fragments are heterocyclic compounds, is practically not touched upon. Similarly, despite the wide choice of catalytic systems currently used in the Negishi reaction, only the most common ones are discussed in the present work. Thus, catalytic systems based on palladium complexes containing ligands of the carbene type have hardly been discussed. When considering the catalysts used in the Negishi reaction, the main attention was paid to catalytic systems based on palladium complexes stabilized by phosphine ligands.

Thus, palladium complexes catalyze the formation of a C–C bond with the participation of aryl halides and nucleophiles (Scheme 1).

ArX + MNu -ArNu + MX

This reaction, first discovered in 1976 by Faurwak, Yutand, Sekiya and Ishikawa using Grignard reagents and organolithium compounds as nucleophiles, was then successfully carried out with the participation of organozin-, aluminum- and organozirconium substrates (Negishi), organotin substrates (Milstein and Steele ), as well as organoboron compounds (Miyaura and Suzuki).

The mechanism of cross-coupling catalyzed by palladium complexes generally includes four main stages. For monodentate phosphine ligands L, the catalytic cycle is shown in Scheme 2.

As an active catalytic particle, it is customary to consider the 14 electron complex of palladium(O), . The first stage of the reaction is the oxidative addition of the aryl halide to with the formation of an α-arylpalladium(II) complex, trans-ArPdXL2 , which is formed after rapid isomerization of the corresponding?///c-complex. The second step in the process is the nucleophilic attack on trans-ArPdXL2, which is called the remetalation step. As a result, a w/?#wc-ArPdnNuL2 complex is formed, in which the palladium(II) atom is bonded to two fragments, Ar and Nu. Next, a trans-r\cis isomerization step is required, since the reductive elimination process, which leads to the cross-coupling reaction product and regeneration of the original palladium complex, occurs exclusively through the formation and subsequent decomposition of the cis-ArPd "NuL2 complex.

When considering palladium catalysts stabilized by monodentate phosphine ligands, and in the case of using relatively low reactive aryl bromides or chlorides as organic electrophiles, the stage that determines the rate of the catalytic cycle is considered to be the oxidative addition process. On the contrary, in the case of using more reactive aryl iodides, it is customary to consider the remetalation step as the rate determining step. The reductive elimination step is also able to determine the rate of the cross-coupling reaction due to the endothermic trans-uis isomerization process.

The study of the sequence of transformations in the study of the mechanism of the cross-coupling reaction is certainly an important task due to the importance of this process for practical chemistry. However, it should be noted that most of the mechanistic studies (for example, those underlying the mechanism presented in Scheme 2) were carried out in isolated systems in which only one of the stages described earlier proceeded, i.e. under conditions rather remotely resembling the catalytic cycle shown in Scheme 2. The general approach underlying the study of the reaction mechanism is to study the elementary steps separately from each other, using as a starting point isolated stable 18-electron complexes, such as the palladium (O) complex Pd°L4 - for oxidative addition, trans- ArPdXL2 - for remetalization and, finally, /??/?a//c-ArPdfINuL2 - for the Ar-Nu formation process. Undoubtedly, the study of individual stages makes it possible to more clearly represent the processes occurring at these individual stages, but this does not provide exhaustive knowledge of the cross-coupling reaction as a whole. Indeed, the study of the reactivity of isolated, and therefore stable, complexes in elementary stages can lead to erroneous results, since a real catalytic cycle can include high-energy and, therefore, unstable complexes that are difficult to detect. For example, it can be noted that anions, cations, and even labile ligands (for example, dba) present in the reaction medium affect the cross-coupling reaction, but these facts cannot be explained within the framework of the reaction mechanism discussed above, which indicates a certain the inferiority of studying the mechanism of the process on the basis of the study of its individual stages.

The efficiency of palladium(O) complexes in the cross-coupling reaction increases in parallel with their ability to activate the Ar-X bond (X = I, Br, C1, OTf) in the oxidative addition reaction. Both stable palladium(O) complexes, for example, and complexes generated in situ from Pd(dba)2 and phosphines are used as catalysts. Palladium(II) complexes, PdX2L2 (X = CI, Br), are also used as palladium(0) precursors. They are reduced either by the nucleophile present in the reaction medium or by a specially added reducing agent if the nucleophile has insufficient reducing power. A mixture of Pd(OAc)2 and phosphines is often used as a source of palladium(0) in the Suzuki reaction. The complexes Pd°L4 and PdChL2 catalyze the formation of the C-C bond in the case of "hard" and "soft" C-nucleophiles. Pd(dba) mixture? and phosphines are more commonly used for "soft" nucleophiles in the Stiehl reaction. Monodentate ligands are effective in cross-coupling reactions involving nucleophiles that are not capable of the p-hydrp elimination process, otherwise the use of bidentate ligands is more effective.

Regardless of the precursor used to obtain palladium(0), the unsaturated 14-electron PdL2 complex is considered as an active species that initiates the catalytic cycle by entering into an oxidative addition reaction (Scheme 2). However, the dependence of reactivity on the method of obtaining PdL2 is often observed. For example, the use of a Pd(PPh3)4 complex as a catalyst is often more efficient than a mixture of Pd(dba)2 with 2 equiv. PPI13. This fact indicates that dba takes part in the catalytic process. It is also postulated that all cross-coupling reactions proceed through the formation of the trap c-ArPdXL2 intermediate during the transmetalation process (Scheme 2). However, some nucleophilic attacks on the m/Jcmc-ArPd^PPh^ complex occur more slowly than the entire catalytic cycle, suggesting a different reaction pathway.

Despite all the shortcomings that are inherent in the study of the mechanism as the sum of individual elementary steps, a more detailed consideration of the mechanism of the cross-coupling reaction will be done in this way, but taking into account all possible substances present in the real reaction mixture, in particular, "labile" ligands , such as dba, anions and cations.

Similar theses in the specialty "Chemistry of organoelement compounds", 02.00.08 VAK code

  • Bismuth(V)Ar3BiX2 Organic Derivatives in Palladium-Catalyzed C-Arylation of Unsaturated Compounds 2008, candidate of chemical sciences Malysheva, Yulia Borisovna

  • Palladium-catalyzed cross-coupling reactions of arylboron compounds with carboxylic acid chlorides. New catalytic systems for the Suzuki reaction 2004, candidate of chemical sciences Korolev, Dmitry Nikolaevich

  • Arylation of ureas and amides with aryl and hetaryl halides under conditions of catalysis by palladium complexes 2004, Candidate of Chemical Sciences Sergeev, Alexey Gennadievich

  • Synthesis of Palladium(II) Complexes with 1,1'-bis(diarylphosphino)metallocenes and Their Electrochemical, Structural, and Catalytic Properties 2003, candidate of chemical sciences Kalsin, Alexander Mikhailovich

  • New methods for modifying steroids using cross-coupling reactions 2006, candidate of chemical sciences Latyshev, Gennady Vladimirovich

Dissertation conclusion on the topic "Chemistry of organoelement compounds", Tsarev, Alexey Alekseevich

substrates

Catalyst

Ni(PPh3)2Cl2 36

It should be noted that, if the combinations of aryl fragments used in the reaction do not contain thermally labile groups, the use of the Suzuki method seems to be more preferable. This is due to the fact that in the case of using arylboronic acids, which have thermal stability, it is possible to carry out the cross-coupling reaction under more severe conditions than in the case of arpzincates, which have greater thermal lability. This makes it possible to obtain sterically loaded products with a high yield, excluding undesirable processes of decomposition of the original organometallic compound. When carrying out the Negishi reaction, in some cases homocoupling products can be observed. This fact can, apparently, be explained by the process of remetalization, which proceeds with copper palladium and organozinc compounds. Interactions of this kind are not characteristic of organoboron compounds.

Using the Negishi reaction, a large number of different biaryls were synthesized, which are interesting from the point of view of biology and medicine. Palladium-catalyzed cross-coupling reactions involving organocyanin compounds were used, for example, to obtain bifenomycin B (biphenomycin B), xenalipin (xenalepin), magnalol (magnalol), (-)-monoterpenylmagnalol ((-)-monoterpenylmagnalol), corupensamine A and B (korupensamine A, B), yupomatsnoida

15 (eupomatenoid-15), cystine (cystine), PDE472, tasosartan (tasosartan) and losartan (losartan) and some other compounds (schemes 43-48).

OH co2n nh2 bifenomycin

Me „magnalol

Me OH corrupensamine A diazonamide A

Me OH corrupensamine B xenalipin

3 stages jupomatenoid-15 co2z co2z

Cbz" catalyst

Z = TMSE clear

Cbz catalyst (% yield): Pd(PfBu3)2 (87), Pd(dba)2/TFP(41), Pd(dba)2/dppf (27)

Pd(dba)2/TFP 73%

CHO diazonamide A multiple stage cystine

V-N precursor of tasosartan N

TBS sec-BuU, TMEDA

THF, -78°С ->

Protocol

Reaction conditions

1. ZnBr2 2. Pd(PPh3)4, THF, Br-> j

1. B(OMe)32. H30+ 3. Pd(PPh3)4, Na2C03, hg-d „ DME, boiling

N VG\ ^ D^DDh.1. TGL "POR

O-™ "o --j:""-O-v

S Me02S"^^ 67% 3"

A, KCH/H ci, PdfPPh, b. 66°C

CI2Pd(PPh3)2, 66°C

2.7. Recent advances in the preparation of biaryls by cross-coupling reaction

In the 2000s, a lot of new works devoted to the study of the cross-coupling reaction appeared. Thus, new catalytic systems have been developed that make it possible to solve such practical problems that could not be solved before. For example, Milne and Buchwald, published in 2004, developed a new phosphine ligand I that allows for the Negishi reaction between various aryl chlorides and organozinc compounds, allowing biaryls with an extremely sterically loaded structure to be obtained in high yield. ligand I

The presence of such groups as CN-, NO2-, NR2~, OR- does not affect the product yield in any way. Tables 12 and 13 present only some of the results obtained.

List of references for dissertation research Candidate of Chemical Sciences Tsarev, Alexey Alekseevich, 2009

1. Time, min Water, % Methanol, %0 30 7015 0 100

2. Time, min Water, % Methanol, %000 20 801500 0 1002500 0 1002501 20 - 80

3. Elemental analysis. Calculated for С10Н9ВУ: С, 53.36; H, 4.03. Found: C, 53.19; H, 3.98.

4. H NMR (CDCb): 5 7.76 (d, J= 7.6 Hz, 1H, 7-H), 7.71 (d, J= 7.6 Hz, 1H, 5-H), 7.28 (t, J= 7.6 Hz, 1Н, 6-Н), 3.36 (dd, J= 17.5 Hz, J= 7.6 Hz, 1Н, 3-Н), 2.70-2.82 (m, 1Н, 2-Н), 2.67 (dd, J= 17.5 Hz, J= 3.8 Hz, 1Н, З"-Н), 1.34 (d, J= 7.3 Hz, ЗН, 2-Me).

5. PS NMR (CDCI3): 5 208.3, 152.9, 138.2, 137.2, 129.0, 122.6, 122.0, 41.8, 35.7, 16.0.

6. Mixture of 4- and 7-bromo-2-methyl-N-indenes (1)

7. Elemental analysis. Calculated for C10H9VP C, 57.44; H, 4.34. Found: C, 57.59;1. H, 4.40.

8. Elemental analysis. Calculated for C10H9CIO: C, 66.49; H, 5.02. Found: C, 66.32; H, 4.95.

9. NMR (CDCb): 5 7.60 (m, IH, 7-H), 7.52 (dd, J= 7.8 Hz, J= 0.9 Hz, 1H, 5-H), 7.29 (m, 1H, 6-H) , 3.35 (m, 1H, 2-H), 2.69 (m, 2H, CH2), 1.30 (d, 3H, Me). 41.3, 33.3, 15.5.

10. A mixture of 4- and 7-chloro-2-methyl-1//-indenes (2)

11. Elemental analysis. Calculated for C10H9CI: C, 72.96; H, 5.51. Found: C, 72.80; H, 5.47.

12. Elemental analysis. Calculated for StsNtsVgO: C, 55.25; H, 4.64. Found: C, 55.35; H, 4.66.1. L17

13. A mixture of 4-bromo-2,5-dimethyl-1//-indene and 7-br(m-2,6-dimethyl-N-1mden (3)

14. Elemental analysis. Calculated for ScNuBr: C, 59.22; H, 4 97. Found: C, 59.35; H, 5.03.

15. Bromo-5-methyl-4,5-dihydro-6/7-cyclopenta6.thiophen-6-one

16. Elemental analysis. Calculated for C\sH7BrOS: C, 41.58; H, 3.05. Found: C, 41.78; H, 3.16.

17. NMR (CDCb): 5 7.77 (s, 1H, 2-H), 3.15 (dd, J= 17.2 Hz, J= 7.0 Hz, 1H, 4-H), 3.04 (m, 1H, 5-H) , 2.50 (dd, J= 17.2 Hz, J= 2.9 Hz, 1H, 4"-H), 1.34 (d, J= 7.5 Hz, 3H, 5-Me).13SNMR (CDCb)" 5 199.3, 165.6, 140.2 , 136.7, 108.4, 47.4, 32.3, 16.7.

18. Bromo-5-methyl-4//-cyclopenta6.thiophene (4)

19. Calculated for C22H22Br2Si: C, 55.71; H, 4.68. Found: C, 56.02; H, 4.77.

20. Bis(4-chloro-2-methyl-1#-nnden-1-yl)(dimethyl)silane (6)

21. Calculated for C22H22Cl2Si: C, 68.56; H, 5.75. Found: C, 68.70; H, 5.88.

22. General procedure for the Negishi reaction involving compounds 5, 7 and 8

23. Compound 9 was prepared according to the general Negishi reaction procedure starting from aryl bromide 5 and phenylmagnesium bromide. Yield 4.54 g (97%) of a white solid, which is an equimolar mixture of rac and meso isomers.

24. Calculated for Cs^Si: C, 87.13; H, 6.88. Found: C, 87.30; H, 6.93.

25. Hs(2,4-d1shetyl-1#-inden-1-yl)(dimethyl)silane (12)

26. Compound 12 was prepared according to the general procedure for the Negishi reaction, starting from aryl bromide 5 and methylmagnesium chloride. Yield 3.34 g (97%) of a white solid, which is an equimolar mixture of rac and meso isomers.

27. Calculated for C24H2sSi: C, 83.66; H, 8.19. Found: C, 83.70; H, 8.26.

28. Compound 13 was prepared according to the general Negishi reaction procedure starting from aryl bromide 5 and 3-trifluoromethylphenylmagnesium bromide. Yield 5.92 g (98%) of a white solid, which is an equimolar mixture of rac and meso isomers.

29. Calculated for C36H3oF6Si: C, 71.50; H, 5.00. Found: C, 71.69; H, 5.13.

30. JPic4-(4-N,N-di^IetnlamIschofshIl)-2-methyl-lH-inden-l-yl.(dimethyl)silane14)

31. Compound 14 was obtained according to the general procedure for the Negishi reaction, starting from aryl bromide 5 and 4-K,.H-dpmetplaminofesh1lmagnesium bromide. Yield 5.10 g (92%) of a white solid, which is an equimolar mixture of paif and meso isomers.

32. Calculated for C38H42N2SK С, 82.26; H, 7.63. Found: C, 82.41; H, 7.58.

33. Calculated for C38H32S2Si: C, 78.57; And, 5.55. Found: C, 78.70; H, 5.46.

34. Compound 16 was prepared according to the general Negishi reaction procedure starting from aryl bromide 5 and 2-trifluoromethylphenylmagnesium bromide. Yield 5.86 g (97%) of a white solid, which is an equimolar mixture of rac- and meso-psomers.

35. Yams4-(4-tert-butylphenyl)-2-metsh|-17/-inden-1-yl(di1methyl)silane (17)

36. Compound 17 was prepared according to the general Negishi reaction procedure, starting from aryl bromide 5 and 4-////7e;/7r-butylfeshmagnesium bromide. Yield 5.70 g (98%) of a white solid, which is a 1:1 mixture of rac and meso isomers.

37. Calculated for C^H^Si: C, 86.84; H, 8.33. Found: C, 86.90; H, 8.39.

38. Compound 18 was prepared according to the general Negishi reaction procedure starting from aryl bromide 7 and phenylmagnesium bromide. Yield 4.72 g (95%) of a white solid, which is an equimolar mixture of rac and meso isomers.

39. b,mc4-(3,5-bis(trifluoromethyl)phenyl)-2,5-dimethyl-1Dr-inden-1-yl(dimethyl)silane (19)

40. Calculated for CsgH^Si: C, 76.97; H, 7.48. Found: C, 77.21; H, 7.56.1. A 23

41. P'c-dimethylsilyl-bisg1=-2-methyl-4-(3-trifluorometh11lfe11yl)inden-1-yl zirconium dichloride (23)

42. Compound 23 was synthesized according to the general procedure starting from ligand "13. An orange solid was obtained in 22% yield.

43. Calculated for CaeH.sCbFeSiZr: С, 56.53; H, 3.69. Found: C, 56.70; H, 3.75.

44. Pc-dimethylsilyl-bisg15-2-1uet11l-4-(4-N,N-dimethylaminophenyl)nnden-1-yl zirconium dichloride (24)

45. Compound 24 was synthesized by the general procedure starting from lpgand 14. An orange solid was obtained in 23% yield.

46. ​​Calculated for C38H4oCl2N2SiZr: C, 63.84; H, 5.64. Found: C, 64.05; II, 5.77.

47. Rc-dimethylsilyl-bis"g|5-2,5-dimethyl-4-phenylinden-1-yl.zirconium dichloride25)

48. Compound 25 was synthesized according to the general procedure starting from ligand 18. An orange solid was obtained in 29% yield.

49. Calculated for C36H34Cl2SiZr: C, 65.83; H, 5.22. Found: C, 65.95; H, 5.31.

50. Compound 26 was synthesized by the general procedure starting from ligand 20. An orange solid was obtained in 25% yield.

51. Calculated for C3oH26Cl2S2SiZr: C, 56.22; H, 4.09. Found: C, 56.41; H, 4.15.

52. Rsh<-диметилсилил-#ис(т15-3-(1-нафтил)-5-метилциклопента6.тиен-6-ил)цирконий дихлорид (27)

53. Compound 27 was synthesized according to the general procedure starting from ligand 22. A red solid was obtained in 22% yield.

54. Calculated for C38H3oCl2S2SiZr: C, 61.59; H, 4.08. Found: C, 61.68; H, 4.15.

55. A mixture of isomeric bis(t/5-2-methyl-4-bromindenyl)zirconium dichlorides (32a and 32b)

56. Elemental analysis. Calculated for C2oHi6Br2Cl2Zr: C, 41.54; H, 2.79. Found: C, 41.69; H, 2.88.

57. JH NMR (CD2C12): isomer 32a, 5 7.54 (d, J= 8.5 Hz, 2H, b^-H), 7.43 (d, J= 7.2 Hz, 2H, 5.5"-H), 7.00 ( dd, J= 8.5 Hz, J= 7.2 Hz, 2H, 7.7"-H), 6.45 (m, 2H, 1,H-H), 6.34 (m, 2H, 3.3"-H), 1.99 (s, 6H, 2.2"-Me).

58. TNMR (CD2C12): isomer 32b, 5 7.57 (d, J= 8.5 Hz, 2H, 6.6"-H), 7.40 (d, J= 7.2 Hz, 2H, 5.5L-H), 6.98 ( dd, J= 8.5 Hz, J- 7.2 Hz, 2H, 7.7^), 6.40 (m, 2H, 1.H-H), 6.36 (m, 2H, 3.3^-H), 2.05 (s , 6H, 2.2"-Me).

59. Elemental analysis. Calculated for CisH2iBrCl2SZr: C, 42.27; H, 4.14. Found: 42.02; And, 4.04.

60. Elemental analysis. Calculated for C22H2oBr2Cl2SiZr: C, 41.65; H, 3.18. Found: C, 41.50; H, 3.11.

61. HilMP (CD2C13): 5 7.60 (dt, J= 8.7 Hz, J= 0.8 Hz, 2Ii, 5.5"-H), 7.52 (dd, J= 7.2 Hz, J= 0.8 Hz, 2H, 7, 7"-H), 6.87 (dd, J= 8.7 Hz, J= 7.2 Hz, 2H, 6.6"-H), 6.83 (m, 2H, 3.3"-H), 2.18 (dia -, J = 0.5 Hz, 6H, 2.2"-Me), 1.26 (s, 6H, SiMe2). 1. Meso-34:

62. Elemental analysis. Calculated for C22H2oBr2Cl2SiZr: C, 41.65; H, 3.18. Found: C, 41.84; H, 3.19.

63. JH NMR (CD2C12): 5 7.57 (d, J= 8.7 Hz, 2H, 5.5"-H), 7.26 (d, J= 7.4 Hz, 2H, 7.7"-H), 6.70 (s , 2H, 3.3"-H), 6.59 (dd, J= 8.7 Hz, J= 7.4 Hz, 2H, 6.6"-H), 2.44 (s, 6H, 2.2"-Me), 1.37 (s, ZN, SiMe), 1.20 (s, ZN, SiMe").

64. Elemental analysis. Calculated for Ci8Hi6Br2Cl2S2SiZr: C, 33.44; H, 2.49. Found: C, 33.47; H, 2.53.

65. Elemental analysis. Calculated for C2oH23CbZr: C, 52.11; H, 5.03. Found: C, 52.34; H, 5.19.

66. Elemental analysis. Calculated for C3H2.Br32r: C, 50.58; H, 2.97. Found: C, 50.62; H, 3.02.

67. Elemental analysis. Calculated for C27H3C^r: C, 62.77; H, 5.85. Found: C, 57.30; H, 5.99.

68. Elemental analysis. Calculated for C26H28Cl2Zr: C, 62.13; H, 5.61. Found: C, 62.34; H, 5.71.

69. Elemental analysis. Calculated for C34H3oCl2SiZr: C, 64.94; H, 4.81. Found: C, 65.08; Н, 4.88.t/5 -2-Methyl-4-p*-tolylindenyl)(775-pentamethylcyclopentadienyl)zirconium dichloride (42)

70. Elemental analysis. Calculated for C27H3oCl2Zr: C, 62.77; H, 5.85. Found: C, 62.95; H, 6.00.

71. Elemental analysis. Calculated for CnH3-^CbXr: C, 63.94; H, 6.29. Found: C, 64.11; H, 6.40.

72. Elemental analysis. Calculated for Cs2Hs2C12r: C, 66.41; H, 5.57. Found: C, 66.67; H, 5.60.

73. Elemental analysis. Calculated for C30H36CI2Z1-: C, 64.49; H, 6.49. Found: C, 64.72; H, 6.62.

74. Elemental analysis. Calculated for C3H3C12r: C, 65.19; H, 5.47. Found: C, 65.53; H, 5.56.

75. NMR (CD2C12): 8 7.10-7.97 (m, YuH, 5,6,7-H in indenyl and naphthyl), 6.22 (dd, J=

76. Elemental analysis. Calculated for C3iH32Cl2Zr: C, 65.70; H, 5.69. Found: C, 65.99; H, 5.85.

77. Elemental analysis. Calculated for C34H32Cl2Zr: C, 67.75; H, 5.35. Found: C, 67.02; H, 5.49.

78. Elemental analysis. Calculated for C^+^ChSZr: C, 56.67; H, 5.15. Found: C, 56.95; H, 5.27.

79. Elemental analysis. Calculated for C24H26Cl2OZr: C, 58.52; H, 5.32. Found: C, 58.66; H, 5.37.

80. Elemental analysis. Calculated for CasHasCbSZr: C, 60.19; H, 5.05. found; C, 60.34; H, 5.20.

81. Elemental analysis. Calculated for Cs2H3C1rOgg: C, 64.84; H, 5.10. Found: : C, 64.70; H, 5.01.

82. Elemental analysis. Calculated for C27H27CI2F3Z1-: C, 56.83; H, 4.77. Found: C, 56.84; H, 4.88

83. Elemental analysis. Calculated for C27H3oCl20Zr: C, 60.88; H, 5.68. Found: C, 61.01; H, 5.75.

84. Elemental analysis. Calculated for C28H33Cl2NZr: C, 61.63; H, 6.10; N, 2.57. Found: C, 61.88; H, 6.24; N, 2.39.

85. NMR (CD2CI2): 5 7.59 (m, 2Н, 2,6-Н in С6Н4), 7.30 (m, 1Н, 7-Н in indenyl), 7.21 (m, 1Н, 5-Н in indenyl), 7.09 (m, 1Н, 6-Н in indenyl), 6.90 (m, 2Н, 3.5-Н in С6Н4), 6.76 (m, 1Н,

86. H in indenyl), 6.22 (m, 1H, 3-H in indenyl), 3.00 (s, 6H, NMe2), 2.19 (s, 3H, 2-Me in indenyl), 2.01 (s, 15H, C. sMes).75.2-Methyl-4-(4-fluorophenyl)indenyl.(75-pentamethylcyclopentadienyl)-zirconium dichloride (58)

87. Elemental analysis. Calculated for C26H27Cl2FZr: C, 59.98; H, 5.23. Found: C, 60.03; H, 5.32.

88. Elemental analysis. Calculated for C28H3oCl202Zr: C, 59.98; H, 5.39. Found: C, 60.11; H, 5.52.

89. Elemental analysis. Calculated for C27H27Cl2NZr: C, 61.46; H, 5.16; N, 2.65. Found: C, . 61.59; H, 5.26; N, 2.49.

90. Elemental analysis. Calculated for C29ll32Cl202Zr: C, 60.61; H, 5.61. Found: C, 60.45; H, 5.77.

91. 1HNMR (CD2C12): 5 8.11 (m, 2H, 3.5-H in SeHC), 7.77 (m, 2H, 2.6-H in SbH), 7.43 (m, 1H, 7-H in indenyl) , 7.30 (dd, J= 7.0 Hz, J= 0.8 Hz, 1Н, 5-Н in indenyl), 7.13 (dd, J= 8.5 Hz,

92. Elemental analysis. Calculated for QjsHjoCbChZr: C, 59.98; H, 5.39. Found: C, 60.18; H, 5.50.

93. Elemental analysis. Calculated for C2.H26C12H £ C, 47.79; H, 4.96. Found: C, 47.87; H, 5.02.

94. H NMR (C6D6): 5 7.02 (m, 1H, 5-H in indenyl), 6.88 (m, 1H, 7-H in indenyl), 6.80 (dd, J= 8.2 Hz, J= 6.8 Hz, 1H , 6-Н in indenyl), 6.45 (m, 1Н, 1-Н in indenyl), 5.56 (d, 2.2

95. Elemental analysis. Calculated for C26H2sCl2Hf: С, 52.94; H, 4.78. Found: C, 53.20; H, 4.89.

96. Elemental analysis. Calculated for CrmH30CHN": C, 53.70; H, 5.01. Found: C, 53.96; H, 5.13.

97. Elemental analysis. Calculated for C3H36CHN £ C, 55.78; H, 5.62. Found: C, 55.91; H, 5.70.

98. Elemental analysis. Calculated for CisHicC^Zr: С, 51.88; H, 4.35. Found: C, 52.10; H, 4.47.

99. Elemental analysis. Calculated for C22H20CI2Z1-: C, 59.18; H, 4.51. Found: C, 59.47; H, 4.68.

100. Using the sequence of actions applied in the case of 41, 500 mg (1.15 mmol) 30, 1.50 ml of a 1.0 M (1.50 mmol) solution of l/-tolylmagnesium chloride in THF, 3.0 ml 0.5

101. M (1.50 mmol) solution of ZnCl2 in THF and 1.15 ml of 0.02 M (0.023 mmol) solution of Pd(P"Bu3)2 in THF lead to the formation of a yellow solid. Yield: 383 mg (75%).

102. Elemental analysis. Calculated for C22H20Cl2Zr: C, 59.18; H, 4.51. Found: C, 59.31; H, 4.60.

103. H NMR (CD2C12): 5 7.05-7.65 (m, 7H, 5,6,7-H in indenyl and 2,4,5,6-H in d/-tolyl), 6.51 (s, 2H, 1 ,3-H in indenyl), 6.02 (s, 5H, C5H5), 2.43 (s, 3H, 3-Me in n*-toll), 2.32 (s, 3H, 2-Me in indenyl).

104. Mixture of isomeric bis(775-2,4-dimethnlindenyl)zirconium dichlorides (72a and 72b)

105. Elemental analysis. Calculated for C22H22Cl2Zr: C, 58.91; H, 4.94. Found: C, 58.99; H, 4.97.

106. NMR (CD2C12): 5 7.23 (m, 2H, 5.5"-Ii), 6.95 (dd, J= 8.1 Hz, J= 6.9 Hz 2H, 6.6"-H), 6.89 (dt, J = 6.9 Hz, J= 1.0 Hz 2H, 7.7x-H), 6.30 (m, 2H, 1,H-H), 6.16 (d, J= 2.2 Hz, 2H, 3.3"-H), 2.39 (s, 6H, 4.4"-H), 2.15 (s, 6H, 2,G-H).

107. Mixture of isomeric bis(775-2-methyl-4-p-tolylindennl)zirconium dichlorondes (73a and 73b)

108. Elemental analysis. Calculated for C34H3oCI2Zr: C, 67.98; H, 5.03. Found: C, 68.11; H, 5.10.

109. Mixture of isomeric bis(g/5-2-methyl-4-p-tolylindenyl)zirconium dichlorides (74a and 74b)

110. Elemental analysis. Calculated for C-wITraChZr: C, 70.15; H, 6.18. Found: C, 70.33; H, 6.25.

111. Elemental analysis. Calculated for Ci9H24Cl2SZr: C, 51.10; H, 5.42. Found: C, 51.22; H, 5.49.

112. Elemental analysis. Calculated for C24H26Cl2SZr: C, 56.67; H, 5.15. Found: C, 56.84; H, 5.23.

113. Elemental analysis. Calculated for C25H28Cl2SZr: C, 57.45; H, 5.40 Found C, 57.57; H, 5.50.

114. Elemental analysis. Calculated for C^s^sCbSZr: C, 57.45; H, 5.40. Found: C, 57.61; H, 5.52.

115. Elemental analysis. Calculated for C^sH^ChSZr: C, 59.55; H, 6.07. Found: C, 59.70; H, 6.16.

116. Ryats-dimethylsilyl-Uns "(/75-2-metnl-4-p-tolylindennl) zirconium dichloride (rac80)

117. Elemental analysis. Calculated for C36H34Cl2SiZr: C, 65.83; H, 5.22. Found: C, 65.94; H, 5.00.

118. Meso-dimethylsilyl-^cis(775-2-methyl-4-p-tolylindenyl)zirconine dichloride (meso-80)

119. Elemental analysis. Calculated for C36H34Cl2SiZr: C, 65.83; H, 5.22. Found: C, 66.14; H, 5.07.

120. Pn(-dimethylsilyl-bis(775-3-(4-tolyl)-5-cyclopeita6.thien-6-yl)zirconium dichloride (81)

121. Elemental analysis. Calculated for C32H3oCl2SSiZr: C, 57.46; H, 4.52. Found: C, 57.70; H, 4.66.

122. Elemental analysis. Calculated for C32H26Cl2Zr: C, 67.11; H, 4.58. Found: C, 67.38; H, 4.65.

123. Elemental analysis. Calculated for C38H3iBr2NZr: C, 60.64; H, 4.15. Found: C, 60.57; H, 4.19.

124. Elemental analysis. Calculated for C34H27Br2NZr: C, 58.29; H, 3.88. Found: C, 58.34; H, 3.92.

125. Rac-dimethylsilyl-bis(2-methyl-4-phenylindenyl-1-yl)zirconium dichloride (85)

126. Elemental analysis. Calculated for Cs+HsoCbSiZr: C, 64.94; H, 4.81. found; C, 65.11; H, 4.92.

127. Zirconium and hafnium complexes containing bromine- and chlorine-substituted rf-cyclopentadienyl ligands of various types were obtained and characterized for the first time, including by X-ray diffraction analysis.

128. It has been shown that the palladium-catalyzed Suzuki-Miyaura reaction using NaBPlu as the arylating agent can be successfully used to synthesize aryl-substituted zirconocenes from the corresponding bromo-substituted substrates.

129. J. F Fauvarque, A. Jutand. Action de divers nucleophiles sur des organopalladiques. // Bull. soc. Chim. fr. 1976, 765.

130. A. Sekiya, N. Ishikawa. The cross-coupling of aryl halides with grignard reagents catalyzed by iodo(phenyl)bis(triphenylphosphine)palladium(II). // J. Organomet. Chem., 1976, 118, 349.

131. E. I. Negishi. Palladium- or Nickel-catalysed cross coupling. A new selective method for C-C bond formation. // acc. Chem. Res., 1982, 15, 340.

132. D. Milstein, J. K. Stille. Palladium-catalyzed coupling of tetraorganotin compounds with aryl and benzyl halides. Synthetic utility and mechanism // J. Am. Chem. Soc., 1979, 101, 4992.

133. N. Miyaura, A. Suzuki. Stereoselective synthesis of arylated (E)-alkenes by the reaction of alk-l-enylboranes with aryl halides in the presence of palladium catalyst. // J. Chem. soc. Chem. Commim., 1979, 866.

134. J. K. Stille. The Palladium-catalyzed cross-coupling Reactions of organotin reagents with organic electrophiles. // Angew. Chem. Int. Ed. Engl., 1986, 25, 508.

135. J. K. Kochi. Organometallic mechanisms and catalysis. // Academic Press, New York, 1978.

136. J. F. Fauvarque, F. Pfluger, M. Troupel. Kinetics of oxidative addition of zerovalent palladium to aromatic iodides. II J. Organomet. Chem., 1981, 208, 419.

137. P. Fit!on, M. P. Johnson, J. E. McKeon. Oxidative additions to palladium(O). // J. Chem. soc. Chem. Commun., 1968, 6.

138. P. Fitton, E. A. Rick. The addition of aryl halides to tetrakis(triphenylphosphine)palladium(0), II J. Organomet. Chem., 1971, 28, 287.

139. A. L. Casado, P. Espinet. On the configuration resulting from oxidative addition of RX to Pd(PPh3)4 and the mechanism of the cis-to-trans isomerization of PdRX(PPh3)2. complexes (R = Aryl, X Halide). // Organomet allies, 1998.17, 954.

140. G. W. Parshall, Sigma-Aryl compounds of nickel, palladium, and platinum. Synthesis and bonding studies. II J. Am. Chem. Soc., 1974, 96, 2360.

141. J. F. Fauvarque, A. Jutand. Arylation of the reformatsky reagent catalyzed by zerovalent complexes of palladium and nickel. II J. Organomet. Chem., 1977, 132, C17.

142. J. F. Fauvarque, A. Jutand. Catalysis of the aiylation of the reformatsky reagent by palladium or nickel complexes. Synthesis of aryl acid esters. and J. Organomet. Chem., 1979, 177, 273.

143. E. Neghishi, T. Takahashi, K. Akiyoshi. Palladium-catalyzed or -promoted reductive carbon-carbon coupling. Effects of phosphines and carbon ligands. // J. Organomet. Chem., 1987, 334, 181.

144. M. S. Driver, J. F. Hartwig. Carbon-nitrogen-bond-forming reductive elimination of arylamines from palladium(II) phosphine complexes. and J. Am. Chem. Soc., 1997, 119, 8232.

145. A. L. Casado, P. Espinet. Mechanism of the Stille Reaction. The Transmetalation step, coupling of Ril and R2SnBu3 catalyzed by trans-PdRiIL2. (Ri = C6C12F3; R2 = Vinyl, 4-Methoxyphenyl; L = AsPh3). // J. Am. Chem. Soc., 1998,120, 8978.

146. A. Gillie, J. K. Stille. Mechanisms of 1,1-reductive elimination from palladium. // J. Am. Chem. Soc., 1980,102, 4933.

147. M. K. Loar, J. K. Stille. Mechanisms of 1,1-reductive elimination from palladium: coupling of styrylmethylpalladium complexes. II J. Arn. Chem. Soc., 1981, 103, 4174.

148. F. Ozawa, T. Ito, Y. Nakamura, A. Yamamoto. Mechanisms of thermal decomposition of trans- and cis-dialkylbis(tertiary phosphine)palladium(II). Reductive elimination and trans to cis isomerization. // Bull. Chem. soc. Jpn., 1981, 54, 1868.

149 G. B. Smith, G. C. Dezeny, D. L. Hughes, A. O. King, T. R. Verhoeven. Mechanistic studies of the Suzuki cross-coupling reaction. II J. Org. Chem., 1994, 59, 8151.

150. V. Farina, B. Krishnan. Large rate accelerations in the stille reaction with tri-2-furylphosphine and triphenylarsine as palladium ligands: mechanistic and synthetic implications. II J. Am. Chem. Soc., 1991, 113, 9585.

151 C. Amatore, F. Pfluger. Mechanism of oxidative addition of palladium(O) with aromatic iodides in toluene, monitored at ultramicroelectrodes. // Orgatiometallics, 1990, 9, 2276.

152. A. Jutand, A. Mosleh. Rate and mechanism of oxidative addition of aryl triflates to zerovalent palladium complexes. Evidence for the formation of cationic (sigma-Aryl) palladium complexes. // Organometallics, 1995, 14, 1810.

153. J. Tsuji. Palladium reagents and catalysts: innovations in organic chemistry. // Wiley, Chichester, 1995.

154 N. Miyaura, A. Suzuki, Palladium-catalyzed cross-coupling reactions of organoboron compounds. II Chem. Rev., 1995, 95, 2457.

155. V. Farina. Transition metal organometallics in organic Synthesis. // Associate Organomet. Chem. II, 1995, 12, 161.

156 J. L. Malleron, J. C. Fiaud, J. Y. Legros. Handbook of palladium-catalyzed organic reactions. Synthetic aspects and catalytic cycles. II Academic Press, New York, 1997.

157 V. Farina, V. ICrishnamurthy, W. J. Scott. The Stille reaction. // Org. React., 1997, 50, 1.

158. H. Geissler (M, Beller, C, Bolm, Eds.). Transition metals for organic synthesis // Wiley-VCH, Weinheim, 1998, 1. 158.

159 F. Henin, J. P. Pete. Synthesis of unsaturated butyrolactones by palladium catalyzed intramolecular carboalkoxylation of homoallylic chloroformates. // Tetrahedron Lett., 1983, 24, 4687.

160. D. Ferroud, J. P. Genet, J. Muzart. Allylic alkylations catalyzed by the couple palladium complexes-alumina. // Tetrahedron Lett., 1984, 25, 4379.

161. B. E. Mann, A. Musco. Phosphorus-31 nuclear magnetic resonance spectroscopic characterization of tertiary phosphine palladium(O) complexes: evidence for 14-electron complexes in solution. and J. Chem. soc. Dalton Trans., 1975, 1673.

162. J. P. Collman, L. S. Hegedus. Principles and applications of organotransition metal chemistry. // Oxford University Press, Oxford, 1980.

163. C.Amatore, A. Jutand, F. Khalil, M. A. M "Barki, L. Mottier. Rates and mechanisms of oxidative addition to zerovalent palladium complexes generated in situ from mixtures of Pd°(dba)2 and triphenylphosphine. // Organometallics, 1993, 12, 3168.

164. J. F. Hartwig, F. Paul. Oxidative addition of aryl bromide after dissociation of phosphine from a two-coordinate palladium(O) complex, Bis(tri-o-tolylphospliine)Palladium(0). // J. Am. Chem. Soc., 1995,117, 5373.

165. S. E. Russell, L. S. Hegedus. Palladium-catalyzed acylation of unsaturated halides by anions of enol ethers. II J. Am. Chem. Soc., 1983,105, 943.

166. W. A. ​​Herrmann, W. R. Thiel, C. BroiBmer, K. Olefe, T. Priermeier, W. Scherer. Dihalogenmethyl)palladium(lI)-complexe aus palladium(0)-vorstufen des dibenzylidenacetons: synthese, strukturchemie und reaktivitatag // J. Organomet. Chem., 1993, 461, 51.

167 C. Amatore, A. Jutand, G. Meyer, H. Atmani, F. Khalil, Ouazzani Chahdi. Comparative reactivity of palladium(O) complexes generated in situ in mixtures of triphenylphosphine or tri-2-furylphosphine and Pd(dba)2. // Organometallics, 1998, 17, 2958.

168. H. A. Dieck, R. F. Heck. Organophosphinepalladium complexes as catalysts for vinylic hydrogen substitution reactions. II J. Am. Chem. Soc., 1974, 96, 1133.

169. C. Amatore, A. Jutand, M. A. M "Barki. Evidence of the formation of zerovalent palladium from Pd (OAc) 2 and triphenylphosphine. // Organometallics, 1992, 11, 3009.

170. F. Ozawa, A. Kobo, T. Hayashi. Generation of tertiary phosphine-coordinated Pd(0) species from Pd(OAc)2 in the catalytic Heck reaction. // Chem. Lett., 1992, 2177.

171. C. Amatore, E. Carre, A. Jutand, M. A. M "Barki. Rates and mechanism of the formation of zerovalent palladium complexes from mixtures of Pd(OAc)2 and tertiary phosphines and their reactivity in oxidative additions. // Organometallics , 1995, 14, 1818.

172 C. Amatore, A. Jutand. Mechanistic and kinetic studies of palladium catalytic systems. I I J. Organomet. Chem., 1999, 576, 254.

173. E. Neghishi, T. Takahashi, K. Akiyoshi. Bis(triphenylphosphine)palladium: its generation, characterization, and reactions. II J. Chem. soc. Chem. Commun., 1986, 1338.

174 C. Amatore, M. Azzabi, A. Jutand. Role and effects of halide ions on the rates and mechanisms of oxidative addition of iodobenzene to low-ligated zerovalent palladium complexes Pd(0)(PPh3)2. II J. Am. Chem. Soc., 1991, 113, 8375.

175 C. Amatore, E. Carre, A. Jutand Evidence for an equilibrium between neutral and cationic arylpalladium(II) complexes in DMF. Mechanism of the reduction of cationic arylpalladium(II) complexes. II Acta Chem. Scand., 1998, 52, 100

176. T. Ishiyama, M. Murata, N. Miyaura. Palladium(0)-catalyzed cross-coupling reaction of alkoxydiboron with haloarenes. a direct procedure for arylboronic esters. II J. Org. Chem, 1995, 60, 7508.

177. A. M. Echavarren, J. K. Stille. Palladium-catalyzed coupling of aryl triflates with organostannanes H J. Am. Chem. Soc., 1987,109, 5478.

178. To Ritter. Synthetic transformations of vinyl and aryl triflates. // Synthesis, 1993, 735.

179. J. Louie, J. F. Hartwig. Transmetalation, involving organotin aryl, thiolate, and amide compounds. An unusual type of dissociative ligand substitution reaction. // J. Am. Chem. Soc., 1995, 117, 11598

180. J. E. Huheey, E. A. Keiter, R L Keitei. Inorganic chemistry: principles of structure and reactivity. // HarperCollins, New York, 1993.11.

181. M. Catellani, G. P. Chiusoli. Palladium-(II) and -(IV) complexes as intermediates in catalytic C-C bond-forming reactions. // J. Organomet. Chem., 1988, 346, C27.

182. M. Alami, F. Ferri, G. Linstrumelle. An Efficient palladium-catalysed reaction of vinyl and aryl halides or triflates with terminal alkynes. // Tetrahedron Lett., 1993, 25, 6403.

183. F. Ozawa, K. Kurihara, M. Fujimori, T. Hidaka, T. Toyoshima, A. Yamamoto. Mechanism of the cross-coupling reaction of phenyl iodide and methylmagnesium iodide catalyzed by trans-PdPh(I)(PEt2Ph)2. // Organomet allies, 1989, 8, 180.

184. J. M. Brown, N. A. Cooley. Observation of stable and transient intermediates in palladium complex-catalysed cross-coupling reactions. II J. Chem. soc. Chem. Commun., 1988, 1345.

185. J. M. Brown, N. A. Cooley. Mapping the reaction pathway in palladium-catalyzed cross-coupling reactions. // Organometallics, 1990, 9, 353

186. M. Portnoy, D. Milstein. Mechanism of aryl chloride oxidative addition to chelated palladium(O) complexes. I I Organometallics, 1993.12, 1665.

187. C. Amatore, A. Jutand. Role of dba in the reactivity of palladium(O) complexes generated in situ from mixtures of Pd(dba)2 and phosphines. // Coord. Chem. Rev., 1998, 511, 178.

188. J. M. Brown, P. J. Guiry. Bite angle dependence of the rate of reductive elimination from diphosphiae palladium complexes. // lnorg. Chim. Acta, 1994, 220, 249.

189. R. A. Widenhoefer, H. A. Zhong, S. T, Buchwald. Direct observation of C~0 reductive elimination from palladium aryl alkoxide complexes to form aryl ethers. // J. Am. Chem. Soc., 1997, 119, 6787.

190. R. A. Widenhoefer, S. T. Buchwald. Electronic dependence of C-0 reductive elimination from palladium (aryl) neopentoxide complexes. II J. Am. Chem. Soc., 1998, 120, 6504.

191. K. Tamao (Eds. B. M. Trost, I. Fleming, G. Pattenden). Comprehensive organic synthesis // Pergamon Press, Oxford, 1991, 3, 819-887.

192. K. Tamao, K. Sumitani, M. Kumada. Selective carbon-carbon bond formation by cross-coupling of Grignard reagents with organic halides. Catalysis by nickel-phosphine complexes // J. Am. Chem. Soc., 1972, 94, 4374.

193. M. Yamamura, I. Moritani, S. I. Murahashi. The reaction of o-vinylpalladium complexes with alkyllithiums. Stereospecific syntheses of olefins from vinyl halides and alkyllithiums. // J. Organomet. Chem., 1975, 91, C39.

194. E. Negishi. Aspects of mechanism and organometallic chemistry (Ed. J. H. Brewster). // Plenum Press, New York, 1978, 285-317.

195. E. Negishi, S. Baba. Novel stereoselective alkenyl-aryl coupling via nickel-catalysed reaction of alkenylanes with aryl halides. // J. Chem. Soc., Chem. Commun., \916, 596b.

196. S. Baba, E. Negishi. A novel stereospecific alkenyl-alkenyl cross-coupling by a palladium-or nickel-catalyzed reaction of alkenylalanes with alkenyl halides. // J. Am. Chem. Soc., 1976, 98, 6729.

197. A. O. King, E. Negishi, F. J. Villani, Jr., A. Silveira, Jr. A general synthesis of terminal and internal arylalkynes by the palladium-catalyzed reaction of alkynylzinc reagents with aryl halides. II J. Org. Chem., 1978, 43, 358.

198. E. Negishi. A genealogy of Pd-catalyzed cross-coupling. II J. Organomet. Chem., 2002, 653, 34.

199. E. Negishi. Organometallics in organic synthesis // Wiley-Interscience, New York, 1980, 532.

200. P. Knochel, J. F. Normant. Addition of functionalized allylic bromides to terminal alkynes. // Tetrahedron Lett., 1984, 25, 1475.

201. P. Knochel, P. Jones (Eds.). Organozinc reagents // Oxford University Press, Oxford, 1999, 354.

202 Y. Gao, K. Harada, T. Hata, H. Urabe, F. Sato. Stereo- and regioselective generation of alkenylzinc reagents via titanium-catalyzed hydrozincation of internal acetylenes. // ./. Org. Chem., 1995, 60, 290.

203. P. Knochel. Metal-catalyzed cross-coupling reactions (Eds. F. Diederich and P. J. Stang) // Wiley-VCH, Weinheim, 1998, 387-419.

204. S. Vettel, A. Vaupel, P. Knochel. Nickel-catalyzed preparations of functionalized organozincs. II J. Org. Chem., 1996, 61.1413.

205. R. F. IIeck. Palladium-catalyzed reactions of organic halides with olefins. // acc. Chem. Res., 1979, 12, 146.

206 E. Negishi, Z. R. Owczarczyk, D. R. Swanson. Strictly regio-controlled method for a-alkenylation of cyclic ketones via palladium-catalyzed cross coupling. // Tetrahedron Lett., 1991, 32, 4453 ().

207. J. Shi, E. Negishi. Pd-catalyzed selective tandem arylation-alkylation of 1,1-dihalo-l-alkenes with aryl- and alkylzinc derivatives to produce a-alkyl-substituted styrene derivatives. // J. Organomet. Chem., 2003, 687, 518.

208 X. Zeng, M. Qian, Q. Ni, E. Negishi. Highly stereoselective synthesis of (£)-2-Methy 1-1,3-dienes by palladium-catalyzed /raws-selective cross-coupling of 1,1-Dibromo-l-alkenes with alkenylzinc reagents. // Angew. Chem., Int. Ed, 2004, 43, 2259.

209 M. R. Netherton, C. Dai, K. Neuschutz, G. C. Fu. Room-temperature alkyl-alkyl Suzuki cross-coupling of alkyl bromides that possess p-hydrogens. UJ. Am. Chem. Soc., 2001, 123, 10099.

210. J. Yin, M. P. Rainka, X.-X. Zhang, S. L. Buchwald. Highly active Suzuki catalyst for the synthesis of sterically hindered biaryls: novel ligand coordination. //./. Am. Chem. Soc., 2002, 124, 1162.

211. R. Giovannini, P. Knochel. Ni(II)-catalyzed cross-coupling between polyfunctional arylzinc derivatives and primary alkyl iodides. II J. Am. Chem. Soc., 1998, 120, 11186.

212. J. Zhou, G. C. Fu. Cross-couplings of unactivated secondary alkyl halides: room-temperature nickel-catalyzed Negishi reactions of alkyl bromides and iodides. II J. Am. Chem. Soc., 2003,125, 14726.

213 C. Dai, G. C. Fu. The first general method for palladium-catalyzed Negishi cross-coupling of aryl and vinyl,chlorides: use of commercially available Pd(P("Bu)3)2 as a catalyst. // J. Am. Chem. Soc., 2001, 123, 2719.

214. J. Zhou, G. C. Fu. Palladium-catalyzed Negishi cross-coupling reactions of unactivated alkyl iodides, bromides, chlorides, and tosylates. II J. Am. Chem. Soc., 2003,125, 12527

215. J. Terao, H. Watanabe, A. Ikumi, H. Kuniyasu, N. Kambe. Nickel-catalyzed cross-coupling reaction of Grignard reagents with alkyl halides and tosylates: remarkable effect of 1,3-butadienes. II J. Am. Chem. Soc., 2002, 124, 4222.

216. W. A. ​​Herrmann, K. Ofele, D. V. Preysing, S. K. Schneider. Phospha-palladacycles and N-heterocyclic carbene palladium complexes: efficient catalysts for C-C-coupling reactions. // J. Organomet. Chem., 2003, 687, 229

217. R. C. Larock. Comprehensive organic transformations: a guide to functional group preparations. // Wiley-VCH New York, 1999, 2, 77-128.

218. G. H. Posner. Substitution reactions using organocopper reagents. // Org. React., 1975, 22, 253.

219. M. F. Semmelhack, P. M. Helquist, L. D. Jones. Synthesis with zerovalent nickel. Coupling of aryl halides with bis(l,5-cyclooctadiene)nickel(0). // J. Am. Chem. Soc., 1971, 93, 5908.

220. R. J. P. Corriu, J. P. Masse. Activation of Grignard reagents by transition-metal complexes. A new and simple synthesis of trans-stilbenes and polyphenyls. // J. Chem. soc. Chem. Commun., 1972, 144a.

221. M. Kumada. Nickel and palladium complex catalyzed cross-coupling reactions of organometallic reagents with organic halides. //Pure Appl. Chem., 1980, 52, 669.

222. E. R. Larson, R. A. Raphael. An improved route to steganone. I I Tetrahedron Lett., 1979, 5041.

223. N. Miyaura, T. Yanagi, A. Suzuki. The palladium-catalyzed cross-coupling reaction of phenylboronic acid with haloarenes in the presence of bases. // synth. Commun., 1981, 11, 513.

224. T. R. Hoye, M. Chen. Studies of palladium-catalyzed cross-coupling reactions for preparation of highly hindered biaryls relevant to the korupensamine/michellamine problem. and J. Org. Chem., 1996, 61, 7940.

225. M. R. Agharahimi, N. A. LeBel. Synthesis of (-)-monoterpenylmagnolol and magnolol. II J. Org. Chem., 1995, 60, 1856.

226. G. P Roth, C. E. Fuller. Palladium cross-coupling reactions of aryl fluorosulfonates: an alternative to triflate chemistry. // J. Org. Chem., 1991, 56, 3493.

227 Y. Okamoto, K. Yoshioka, T. Yamana, H. Moil. Palladium-catalyzed cross-coupling of bromobenzenes, containing an acetyl or a formyl group, with organozinc reagents. // J. Organomet. Chem., 1989, 369, 285.

228. M. Rottlander, N. Palmer, P. Knochcl, Selective Pd(0)-catalyzed arylations with new electrophilic or nucleophilic multi-coupling reagents. // Synlett, 1996, 573.

229 C. A. Quesnelle, O. B. Familoni, V. Snieckus. Directed ortho metalation cross coupling connections. Nickel (0) - catalyzed cross coupling of aryl triflates with organozinc reagents. // Synlett, 1994, 349.

230. T. Ohe, N. Miyaura, A. Suzuki. Palladium-catalyzed cross-coupling reaction of organoboron compounds with organic triflates. II J. Org. Chem., 1993, 58, 2201.

231. V. Aranyos, A. M. Castnao, H. Grennberg. An application of the Stille coupling for the preparation of arylated phthalonitriles and phthalocyanines. II Acta Chem. Scand., 1999, 53, 714.

232 K. Koch, R. J. Chambers, M. S. Biggers. A direct synthesis of pharmacologically active o/Y/josubstituted biaryls: a combined directed metalation-palladium catalyzed cross coupling approach using aryl oxazolines or benzamides. // Synlett, 1994, 347.

233. S. Saito, S. Oh-tani, N. Miyaura. Synthesis of biaryls via a nickel(0)-catalyzed cross-coupling reaction of chloroarenes with arylboronic acids. // J.Org. Chem., 1997, 62, 8024.

234. J. A. Miller, R. P. Farrell. Preparation of unsymmetrical biaryls via Ni- or Pd-catalyzed coupling of aryl chlorides with arylzincs. // Tetrahedron Lett., 1998, 39, 6441.

235. J. Huang, S. P. Nolan. Efficient cross-coupling of aryl chlorides with aryl Grignard reagents (Kumada reaction) mediated by a palladium/imidazolium chloride system. //./. Am. Chem. Soc., 1999,121, 9889.

236. J. Galland, M. Savignac, J. Genet. Cross-coupling of chloroarenes with boronic acids using a water-soluble nickel catalyst. // Tetrahedron Lett., 1999, 40, 2323.

237. K. Takagi. Ultrasound-promoted synthesis of arylzinc compounds using zinc powder and their application to palladium(0)-catalyzed synthesis of multifunctional biaryls. // Chem. Lett, 1993, 469.

238.E.I. Negishi, T. Takahashi, A. O. King. Synthesis of biaryls via palladium-catalyzed cross-coupling 2-methyl-4" nitrobiphenyl. // Org. Synth., 1988, 66, 67;

239. A. Palmgren, A. Thorarensen, J. Beckvall. Efficient synthesis of symmetrical 2,5-disubstituted benzoquinones via palladium-catalyzed double Negishi coupling. // J.Org. Chem., 1998, 63, 3764.

240. K. Manabe, K. Okamura, T. Date, K. Koga. Receptors for oxo acids: effects of intra-ion-pair hydrogen bonding on acid-base equilibria. // J.Org. Chem., 1993, 58, 6692.

241. J. C. Adrian, Jr., C. S. Wilcox. Chemistry of synthetic receptors and functional group arrays. 10. Orderly functional group dyads. Recognition of biotin and adenine derivatives by a new synthetic host. II J. Am. Chem. Soc., 1989, 111, 8055.

242. S. Coleman, E. B. Grant. Application of a Cu(I)-mediated biaryl cross-coupling reaction to the synthesis of oxygenated 1, G-binaphthalenes. // Tetrahedron Lett., 1993, 34, 2225.

243 U. Schmidt, R. Meyer, V. Leitenberger, H. Griesser, A. Lieberknecht. Total synthesis of the biphenomycins; synthesis of biphenomycin B. // Synthesis, 1992, 1025.

244. T. Bach, M. Bartels. 2,3-Disubstituted and 2,3,5-trisubstituted benzofurans by regioselective Pd-catalyzed cross-coupling reactions; a short synthesis of eupomatenoid-15. // Synlett, 2001, 1284.

245 P. Nshimyumukiza, D. Cahard, J. Rouden, M. C. Lasne, J. C. Plaquevent. Construction of functionalized/substituted bipyridines by means of Negishi cross-coupling reactions. Formal synthesis of (±)-cytisine. // Tetrahedron Lett., 2001, 42, 7787.

246. P. W. Manley, M. Acemoglu, W. Marterer, W. Pachinger. Large-scale Negishi coupling as applied to the synthesis of PDE472, an inhibitor of phosphodiesterase type 4D. // Org. Process Res. Dev., 2003, 7, 436.

247. W. Cabri, R. D. Fabio. From bench to market: the evolution of chemical synthesis. // Oxford University Press, 2000, 6, 120-145.

248. K. S. Feldman, K. J. Eastman, G. Lessene. Diazonamide synthesis studies: use of Negishi coupling to fashion diazonamide-related biaryls with defined axial chirality. // Org. Lett., 2002, 4, 3525.

249 M. R. Reeder, H. E. Gleaves, S. A. Hoover, R. J. Imbordino, J. J. Pangborn. An improved method for the palladium cross-coupling reaction of oxazol-2-ylzinc derivatives with aryl bromides. // Org. Process Res. Dev., 2003, 7, 696.

250. T. Bach, S. Heuser. Synthesis of 2"-substituted 4-bromo-2.4"-bithiazoles by regioselective cross-coupling reactions. // J. Org. Chem., 2002, 67, 5789.

251. J. E. Milne, S. L. Buchwald. An extremly active catalyst for the Negishi cross-coupling reaction. II J. Am. Chem. Soc., 2004,126, 13028.

252 G. Manolikakes, M. A. Schade, C. M. Hernandez, H. Mayr, P. Knochcl. Negishi cross-couplings of unsatureted halides bearing relatively acidic hydrogen atoms with organozinc reagents. // Org. Lett., 2008, 10, 2765.

253. M. Ohff, A. OhfF, D. Milstein. Highly active Pdll cyclomelallated imine catalysts for the Heck reaction. // Chemical Communications, 1999, 4, 357.

254 K. Nikitin, H. Mueller-Bunz, Y. Ortin, M. J. McGlinchey. Joining the rings: the preparation of 2- and 3-indenyl-triptycenes, and curious related processes. // Organic & Biomo/ecular Chemistry. 2007, 5, 1952.

255 Y. Hatanaka, K. Goda, T. Hiyama, a-Selective cross-coupling reaction of allyltrifluorosilanes: remarkable ligand effect on the regiochemistry, Tetrahedron Lett., 1994, 35, 6511.

256. Y. Hatanaka, Y. Ebina, T. Hiyama, a-Selective cross-coupling reaction of allyltrifluorosilanes: a new approach to regiochemical control in allylic systems. // J. Am. Chem. Soc., 1991, 113, 7075.

257. S. Patai, Z. Rappopoit. The Chemistry of Organic Silicon Compounds // Wiley, 1989.

258. M.-C. Otto, G. Salo. Thiophene analogs of indenes. I. Synthesis of indanone analogs. // Acta Chemica Scandinavica, 1966, 20, 1577.

259 J. Frohlich. Halogen dance reactions at thiophenes and furans: selective access to a variety of new trisubstituted derivatives. // Bull. soc. Chim. Beiges. 1996, 105, 615.

260. E. Negishi, X. Zeng, Z. Tan, M. Qian, Q. Hu, Z. Huang, Metal-Catalyzed Cross-Coupling Reactions (2nd Edition). // Oxford, 2004, 2, 815.

261. J. Hassan, M. Sevignon, C. Gozzi, E. Schulz, M. Lemaire, Aryl-Aryl bond formation one" century after the discovery of the Ullmann reaction. // Chem. Rev. 2002, 102, 1359.

262 D. J. Cardin, M. F. Lappert, C. L. Raston, Chemistry of organo-zirconium and -hafnium compounds. //Raston/EllisHonwoodLtd., 1986.

263 E. F. Abel, F. G. A. Stone, G. Wilkinson, Comprehensive Organometallic Chemistry II. // Pergamort, 1995, 4.

264 R. H. Crabtree, D. M. P. Mingos, Comprehensive Organometallic Chemistry III. // Elsevier, 2007, 4.

265. II. H. Brintzinger, D. Fischer, R. Miilhaupt, B. Rieger, R. M. Waymouth. Stereospecific olefin polymerization with chiral metallocene catalysts. // Angew. Chem., Int. Ed., 1995, 34, 1143.

266. G. W. Coates, R. M. Waymouth. Oscillating stereocontrol: a strategy for the synthesis of thermoplastic elastomeric pplypropylene // Science, 1995, 267, 217.

267. E. Hauptman, R. M. Waymouth, J. M. Ziller. Stereoblock polypropylene: ligand effects on the stereospecificity of 2-arylindene zirconocene catalysts. // J. Am. Chem. Soc., 1995, 117, 11586.

268 X. Zhang, Q. Zhu, I. A. Guzei, R. F. Jordan. General synthesis of racemic Me2Si-bridgcd bis(indenyl) zirconocene complexes. // J. Am. Chem. Soc., 2000, 122, 8093.

269. R. W Lin, T.E. DeSoto, J. F. Balhoff. Zirconocene isomerization process. // U.S. Pat. Appl. PubL, 1998, 005780660.

270. R W. Lin. Catalytic process for isomerizing metallocenes. II U.S. Pat. Appl. PubL, 1998, 005965759.

271. G. G. Hlatky. Heterogeneous single-site catalysts for olefin polymerization. II Chem. Rev. 2000, 100, 1347.

272. P. Knochel, A. Krasovskiy, I. Sapountzis. Handbook of Functionalized Organometallics: Applications in Synthesis. // Wiley-VCH, 2005.

273. R. D. Rieke. The preparation of highly reactive metals and development of novel organometallic reagents. // Aldrichimica Acta, 2000, 33, 52

274 S. Sase, M. Jaric, A. Metzger, V. Malakhov, P. Knochel. One-Pot Negishi cross-coupling reactions of in situ generated zinc reagents with aryl chlorides, bromides, and triflates. // J. Org. Chem., 2008, 73, 7380.

275. R. M. Buck, N. Vinayavekhin, R. F. Jordan. Control of ansa-zirconocene stereochemistry by reversible exchange of cyclopentadienyl and chloride ligands. // J. Am. Chem. Soc., 2007, 129, 3468.

276 B. E. Bosch, I. Briimer, K. Kunz, G. Erker, R. Frohlich, S. Kotila. Structural characterization of heterodimetallic Zr/Pd and Zr/Rh catalyst precursors containing the C5H4PPh2 ligand. // Organometallics, 2000, 19, 1255.

277. G. M. Sosnovskii, A. P. Lugovskii, and I. G. Tishchenko. Synthesis of meso-substituted tricarbocyanine dyes with an o-phenylene bridge in the chromophore. // Z. Org. Khim. 1983, 19, 2143.

278. I. E. Nifant "ev, A. A. Sitnikov, N. V. Andriukhova, I. P. Laishevtsev, Y. N. Luzikov, A facile synthesis of 2-aryIndenes by Pd-catalyzed direct arylation of indene with aryl iodides. // Tetrahedron Letters 2002, 43, 3213.

Please note that the scientific texts presented above are posted for review and obtained through original dissertation text recognition (OCR). In this connection, they may contain errors related to the imperfection of recognition algorithms. There are no such errors in the PDF files of dissertations and abstracts that we deliver.